preparation and characterization of biopolymer compounds

141
Preparation and characterization of biopolymer compounds containing poly-3-hydroxyalkanoates and polylactic acid by Manoj Nerkar A thesis submitted to the Department of Chemical Engineering In conformity with the requirements for the degree of Doctor of Philosophy Queen’s University, Kingston, Ontario, Canada (September, 2014) Copyright © Manoj Nerkar, 2014

Upload: doanxuyen

Post on 13-Feb-2017

231 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Preparation and characterization of biopolymer compounds

Preparation and characterization of biopolymer compounds

containing poly-3-hydroxyalkanoates and polylactic acid

by

Manoj Nerkar

A thesis submitted to the Department of Chemical Engineering

In conformity with the requirements for the degree of

Doctor of Philosophy

Queen’s University, Kingston, Ontario, Canada

(September, 2014)

Copyright © Manoj Nerkar, 2014

Page 2: Preparation and characterization of biopolymer compounds

i

Abstract

This thesis is focused on developing cost effective and environmentally friendly techniques to

improve the properties and processability of biopolyesters through compounding and reactive

modification. Specifically, elastomeric medium-chain-length poly(3-hydroxyalkanoates) (MCL

PHA) have been evaluated as potential impact modifiers for poly(lactic acid) (PLA) and poly-3-

hydroxybutyrate (PHB), using conventional melt compounding. The Mark-Houwink constants,

absolute molecular weight distributions and the absolute molecular weight (MW) averages of

MCL PHAs with predominantly 3-hydroxyoctanoate (PHO), 3-hydroxynonanoate (PHN) or 3-

hydroxydodecanoate (PHDD) content were determined and ranged between 18,200 for PHN to

172,000 Da for PHO. Detailed thermal and rheological characterization revealed that PHO had

the highest viscosity, and was thus the best candidate as impact modifier for PHB and PLA. Melt

compounded PHB/PHO and PLA/PHO blends showed improved tensile strain at break and

unnotched impact strength upon addition of up to 30 wt.% PHO in PHB and 15 wt.% PHO in

PLA. This was counteracted by decreased Young’s modulus due to lower blend crystallinity. The

droplet-matrix morphology coarsened as PHO content increased beyond 5 wt.%, due to PHO

coalescence attributed to viscosity mismatch between blend components. PHO was reacted

using lauroyl peroxide to increase its viscosity through partial cross-linking, thus improving the

morphology but the mechanical properties showed only moderate improvements, presumably

due to high PHO gel content which compromised its elastomeric nature. Reactive compounding

by radical mediated solvent-free grafting of triallyl trimesate (TAM) coagent was employed to

improve blend properties. Reactively modified PLA had higher molar mass, melt viscosity and

enhanced strain hardening. Additionally it showed a distinct crystallization peak upon cooling

Page 3: Preparation and characterization of biopolymer compounds

ii

with disappearance of the cold crystallization peak, indicative of a nucleation effect. PLA

modified using a multi-functional epoxide oligomeric chain extender yielded similar

improvements in rheological properties, but no considerable change in crystallization. This

coagent modification approach also increased the viscosity of PHO, and improved both

extrudate appearance and handling. Coagent modified PLA/PHO blends demonstrated

significant improvement in crystallization and rheological properties, similar to those seen in

the coagent modified PLA alone, while the mechanical properties remained unaffected.

Page 4: Preparation and characterization of biopolymer compounds

iii

Co-Authorship

This thesis contains seven chapters that present results that have been published in the form of

original journal articles as well as material that is in preparation for submission. The complete

citations for the published papers and chapter in which they appear are provided bellow:

Chapter 3: Nerkar M, Ramsay J, Ramsay B, Kontopoulou M, Hutchinson R.

Determination of Mark-Houwink parameters and absolute molecular weight of medium-

chain-length poly(3-hydroxyalkanoates). Journal of Polymers and the Environment 2013

(21): 24-29

Chapter 4: Nerkar M, Ramsay J, Ramsay B, Kontopoulou M. Melt compounded

blends of short and medium-chain-length poly-3-hydroxyalkanoates. Journal of

Polymers and the Environment 2014 (22): 236-243

Chapter 5 : Nerkar M, Ramsay J, Ramsay B, Kontopoulou M. Dramatic Improvements

in Strain Hardening and Crystallization kinetics of PLA by simple reactive modification in

the melt state. Macromolecular Materials and Engineering (Accepted –May 2014)

Chapter 6 : Nerkar M, Ramsay J, Ramsay B, Kontopoulou M. Improvements in the

extensional rheology, thermal properties and morphology of poly(lactic acid)/ poly-3-

hydroxyoctanoate blends through reactive modification (To be submitted)

All the papers and manuscripts were co-authored and reviewed prior to publication by

Professors Marianna Kontopoulou, Juliana A. Ramsay and Bruce A. Ramsay. The first paper

(chapter 3) was co-authored by Professor Robin Hutchinson who directed the experiments to

determine true molecular weight of medium-chain-length polyhydroxyalkanoate. Dr.

Alexandros Vasileiou helped in the synthesis and characterization of epoxidized

Page 5: Preparation and characterization of biopolymer compounds

iv

polyhydroxyoctanoate presented in Appendix A. Hang Li performed fermentations that made

medium-chain-length polyhydroxyalkanoates. All the rest of the experimental work and

manuscript preparation were performed by the author of this thesis.

Page 6: Preparation and characterization of biopolymer compounds

v

Acknowledgements

It would not have been possible to complete my doctoral thesis without the support and help

from my supervisors, staff members, lab mates, family and friends.

I consider myself lucky to have a supervisor like Professor Marianna Kontopoulou. I appreciate

her for being a great support to me and to my family in a foreign land. Apart from her endless

technical support and research guidance, she was always motivating and encouraging. She has

been a wonderful person to work with. It’s my great privilege to be co-supervised by Professor

Juliana A Ramsay. I would like to thank her for her technical guidance and for taking time to

critically reviewing all my work. I was privileged to have the opportunity to work closely with

Dr. Bruce A Ramsay. He was instrumental in providing his expert opinion about biopolymers. I

would like to thank him, along with his team members, Dr. Zhiyong Sun, Hang Li and Eric Potter

for providing the polymers required for my research.

Professor Robin Hutchinson is greatly acknowledged for his support in molecular weight

characterization of polymers. Timely support from Dr. Kalam Mir, Steven Hodgson and Kelly

Sedore is highly appreciated. Charlie Cooney from the materials department provided judicious

help in conducting SEM analysis of samples.

I would like to thank my lab mates Osayuki Osazuwa, Ying Zhang, Andrew Powell, Praphulla and

Dr. Alexandros Vasileiou for their friendship, technical discussions and help.

Funding support from the Natural Science and Engineering Research Council (NSERC), Queen's

University Graduate Award and Xerox Research Center of Canada is gratefully acknowledged.

I take this opportunity to thank my parents Mrs. Rajani Gokul Nerkar, Mr. Gokul

Madhavrao Nerkar, and my in-laws, Mrs. Shailaja Arvind Sonar and the late Mr. Arvind

Page 7: Preparation and characterization of biopolymer compounds

vi

Lilachand Sonar and all family members including brothers, brother-in-laws and sister-in-laws

and friends from around the world who provided all kinds of support throughout my life.

Poonam, my lovely wife, is the person who inspired me to get doctorate. I cannot thank her

enough for all of her sacrifices and for being with me all the time. She took care of our kids and

their illness single handedly when I was busy studying for exams, attending conferences or

writing manuscripts.

My acknowledgement cannot be completed without mentioning my adorable son Aayush and

my beautiful daughter Anishka. I thank them for their unconditional love.

Page 8: Preparation and characterization of biopolymer compounds

vii

Contents

Abstract ...................................................................................................................................................... i

Co-Authorship .......................................................................................................................................... iii

Acknowledgement .................................................................................................................................... v

List of Figures ............................................................................................................................................ x

List of Tables ...........................................................................................................................................xiii

List of Schemes........................................................................................................................................ xiv

Nomenclature ......................................................................................................................................... xv

Abbreviations .......................................................................................................................................... xvi

Chapter 1 Introduction .............................................................................................................................. 1

1.1 Thesis objectives ........................................................................................................................... 2

1.1.1 Characterization of MCL PHA ................................................................................................ 3

1.1.2 Blends of MCL PHA with brittle biopolymers ........................................................................ 3

1.1.3 Enhancement of melt viscosity of MCL PHA for blending with PHB and PLA ....................... 3

1.1.4 Improved melt strength and crystallization of PLA and its blends with MCL PHA by

reactive modification ............................................................................................................................ 4

1.2 Thesis organization ....................................................................................................................... 4

Chapter 2 Literature Review ...................................................................................................................... 6

2.1 Polyhydroxyalkanoates ................................................................................................................. 6

Medium-chain-length (MCL) PHA ......................................................................................................... 7

2.1.1 Impact modification of PHB .................................................................................................. 8

2.1.2 Plasticization ......................................................................................................................... 9

2.1.3 Nucleation ........................................................................................................................... 10

2.1.4 Chain extension ................................................................................................................... 12

2.2 Polylactic acid (PLA) .................................................................................................................... 12

2.2.1 Impact modification ............................................................................................................ 14

2.2.2 Plasticization ....................................................................................................................... 14

2.2.3 Nucleation ........................................................................................................................... 15

2.2.4 Conditioning ........................................................................................................................ 15

2.2.5 Chain extension ................................................................................................................... 16

2.2.6 Epoxy based chain extension .............................................................................................. 16

Page 9: Preparation and characterization of biopolymer compounds

viii

2.2.7 Peroxide-mediated cross-linking......................................................................................... 18

2.2.8 Multifunctional coagents .................................................................................................... 19

Chapter 3 Determination of Mark-Houwink parameters and absolute molecular weight of medium-

chain-length poly(3- hydroxyalkanoates)* ................................................................................................. 21

3.1 Introduction ................................................................................................................................ 21

3.2 Experimental ............................................................................................................................... 23

3.2.1 Materials ............................................................................................................................. 23

3.2.2 Methods .............................................................................................................................. 24

3.3 Results and Discussion ................................................................................................................ 26

3.3.1 Molecular weight determination ........................................................................................ 26

3.3.2 Thermal and rheological characterization .......................................................................... 32

3.4 Conclusion ................................................................................................................................... 34

Chapter 4 Melt compounded blends of short and medium-chain-length poly-3-hydroxyalkanoates* .. 35

4.1 Introduction ................................................................................................................................ 35

4.2 Experimental ............................................................................................................................... 37

4.2.1 Materials ............................................................................................................................. 37

4.2.2 Compounding ...................................................................................................................... 37

4.2.3 PHO cross-linking ................................................................................................................ 37

4.2.4 Blend Characterization ........................................................................................................ 38

4.3 Results and discussion ................................................................................................................ 40

4.3.1 Thermal and rheological properties .................................................................................... 40

4.3.2 Morphology ......................................................................................................................... 44

4.3.3 Mechanical properties ........................................................................................................ 46

4.3.4 Cross-linking of PHO ............................................................................................................ 47

4.4 Conclusions ................................................................................................................................. 52

Chapter 5 Dramatic improvements in strain hardening and crystallization kinetics of PLA by simple

reactive modification in the melt state*..................................................................................................... 54

5.1 Introduction ................................................................................................................................ 54

5.2 Experimental ............................................................................................................................... 56

5.3 Results and Discussion ................................................................................................................ 59

5.3.1 Rheological characterization ............................................................................................... 59

5.3.2 Thermal properties ............................................................................................................. 63

Page 10: Preparation and characterization of biopolymer compounds

ix

5.4 Conclusions ................................................................................................................................. 66

Chapter 6 Improvements in the extensional rheology, thermal properties and morphology of

poly(lactic acid)/ poly-3-hydroxyoctanoate blends through reactive modification ................................... 68

6.1 Introduction ................................................................................................................................ 68

6.2 Experimental ............................................................................................................................... 70

6.2.1 Materials ............................................................................................................................. 70

6.2.2 Compounding ...................................................................................................................... 70

6.2.3 Characterization .................................................................................................................. 71

6.3 Results and Discussion ................................................................................................................ 74

6.3.1 Blends of PLA with PHO ...................................................................................................... 74

6.3.2 Reactive modification of PHO ............................................................................................. 77

6.3.3 Reactive modification of PLA .............................................................................................. 81

6.3.4 Reactive compounding of PLA with PHO ............................................................................ 85

6.3.5 Thermal and rheological properties .................................................................................... 86

6.3.6 Blend morphology ............................................................................................................... 86

6.3.7 Mechanical properties ........................................................................................................ 89

6.4 Conclusions ................................................................................................................................. 91

Chapter 7 Thesis overview ....................................................................................................................... 92

7.1 Thesis overview ........................................................................................................................... 92

7.2 Conclusions ................................................................................................................................. 93

7.3 Significant contributions ............................................................................................................. 96

7.4 Recommendation for future work .............................................................................................. 97

References .................................................................................................................................................. 99

Appendix A - Improved viscosity ratio and compatibility of poly (lactic acid) and polyhydroxyoctanoate

blends ........................................................................................................................................................ 114

Page 11: Preparation and characterization of biopolymer compounds

x

List of Figures

Figure 3.1 a) Experimental Mark-Houwink data for PHO (three replicates), b) best fit Mark-

Houwink relationships for all copolymer samples. ....................................................................... 28

Figure 3.2 Molecular weight distributions for PHO (three replicates) as measured by a)

Waters/Wyatt SEC, light-scattering detector; b) Waters/Wyatt SEC, universal calibration; c)

Viscotek SEC, triple detection; and d) Viscotek –SEC, universal calibration. ............................... 30

Figure 3.3 Molecular weight distributions of four MCL PHA polymers (Viscotek SEC, universal

calibration). ................................................................................................................................... 32

Figure 3.4 Viscosity as a function of molecular weight ................................................................ 33

Figure 4.1 TGA curves for PHO, PHB and the 85/15 PHB/PHO blend at 190 °C ........................... 41

Figure 4.2 a) DSC endotherm (2nd heating cycle) and b) DSC exotherm of PHB and PHB/PHO

blends ............................................................................................................................................ 43

Figure 4.3 Rheological properties of PHB and PHO and effect of peroxide cross-linking on a)

complex viscosity, b) storage modulus and c) loss tangent, tan δ, measured at 190 °C ............. 45

Figure 4.4 Scanning electron microscopy blends containing a) 5 wt.%, b) 10 wt.%, c) 15 wt.%, d)

20 wt.% and e) 30 wt.% of PHO at 2000x magnification .............................................................. 46

Figure 4.5 a) Un-notched impact strength and tensile strain b) tensile stress and Young’s

modulus of pristine PHB and its blends ........................................................................................ 48

Figure 4.6 Cure curves of PHO with 0.1 wt.% lauroyl peroxide as a function of temperature at a

frequency of 1 Hz. ......................................................................................................................... 49

Page 12: Preparation and characterization of biopolymer compounds

xi

Figure 4.7 SEM of blends of PHB/cross-linked PHO 70/30 blends a) uncross-linked (viscosity

ratio, λ=0.03) b) cross-linked with 0.06 wt.% of lauroyl peroxide (λ=0.15) c) cross-linked with 0.2

wt.% of lauroyl peroxide (λ=0.36) d) cross-linked with 0.5 wt.% of lauroyl peroxide (λ=3.73). .. 51

Figure 5.1 a) Complex viscosity as a function of frequency and b) phase degree as a function of

complex modulus at 180 °C. ......................................................................................................... 60

Figure 5.2 Tensile stress growth coefficient (ηE+) of TAM and GMA modified PLA as a function of

strain rate and time at Hencky strain rates of 0.1, 1 and 10 s-1 at 180 °C. Curves are shifted by an

arbitrary factor for the sake of clarity. Solid lines represent the LVE envelop (3E+) for each

sample. .......................................................................................................................................... 62

Figure 5.3 DSC a) 2nd heating scan at rate of 5 °C/min b) cooling scan at the rate of 5 °C/min . 63

Figure 5.4 Relative degree of crystallinity as a function of time a) isothermal crystallization

experiments; (-) PLA/TAM at 135 °C, ()PLA/TAM at 140 °C, ()PLA/TAM at 150 °C, (◆)

PLA/GMA at 135 °C and (b) non-isothermal crystallization experiments; ()PLA/TAM at 2.5

°C/min, (◆)PLA/TAM at 5 °C/min, (o)PLA/TAM at 20 °C/min, ()PLA/GMA at 2.5 °C/min,

()PLA/GMA at 5 °C/min .............................................................................................................. 65

Figure 6.1 Scanning electron microscope images of PLA blend containing a) 5 wt.%, b) 10 wt.%,

c) 15 wt.% and d) 20 wt.% of PHO. ............................................................................................... 75

Figure 6.2 Effect of TAM content on the rheological properties of PHO with DCP content

remaining constant a) Complex viscosity b) storage modulus and c) tan δ ................................. 78

Figure 6.3 a) unmodified PHO after extrusion b) PHO/0.3/1 after extrusion .............................. 79

Figure 6.4 Effect of DCP amount on a) Complex viscosity b) storage modulus and c) tan δ of

coagent modified PHO (PHO 0.3/1 and PHO 0.5/1 yielded 23 and 42 % gel respectively) ....... 80

Page 13: Preparation and characterization of biopolymer compounds

xii

Figure 6.5 Effect of coagent modification on the complex viscosity of PLA and PHO ................. 81

Figure 6.6 Tensile stress growth coefficient (ηE+) of PLA/0.3/1 and (PLA/PHO)/0.3/1 as a

function of strain rate and time at Hencky strain rates of 0.1, 1 and 10 s-1 at 180 °C. Curves are

shifted by an arbitrary factor for the sake of clarity. Dotted lines represent the LVE envelop for

each sample. ................................................................................................................................. 82

Figure 6.7 DSC (a) cooling exotherm (b) heating endotherm of coagent-modified PLA and

PLA/PHO blends ............................................................................................................................ 84

Figure 6.8 Hot stage microscopy of a) PLA, b) PLA/0.3/1 at 135 °C ............................................. 85

Figure 6.9 Effect of DCP and TAM on a) complex viscosity b) storage modulus and c) tan δ of

PLA/PHO blends ............................................................................................................................ 87

Figure 6.10 Scanning electron microscopy of PLA/PHO (90/10) blend a) unmodified b)

(PLA/PHO)/0.5 c) (PLA/PHO)/1 ..................................................................................................... 88

Figure 6.11 Effect of coagent modification on morphology of PLA/PHO blends a) 95/05 b) 90/10

c) 80/20 (wt./wt.%) (samples reacted with coagent were not etched); Top raw without coagent;

bottom raw with coagent ............................................................................................................. 89

Page 14: Preparation and characterization of biopolymer compounds

xiii

List of Tables

Table 2.1 Properties of PLA [37] ................................................................................................... 13

Table 3.1 Mark Houwink calibration constants from best fit to triple detector analysis of MCL

PHA samples in THF ([Ƞ]=K(MW)a). .............................................................................................. 29

Table 3.2 Number-average (Mn), weight-average (Mw), and dispersity (PDI) values determined

from SEC analysis of MCL PHA samples. ....................................................................................... 31

Table 3.3 Thermal properties of MCL PHAs .................................................................................. 33

Table 4.1 Crystallization temperature (TC), first and second melting peaks (TM1 and TM2

respectively), % crystallinity and degradation onset temperature for PHB and PHB/PHO blends.

....................................................................................................................................................... 42

Table 4.2 Comparison of mechanical properties of PHB/PHO 70/30 blends containing uncross-

linked and cross-linked PHO ......................................................................................................... 52

Table 5.1 Material characterization .............................................................................................. 61

Table 5.2 Isothermal Avrami constants and crystallization half time for PLA/GMA and PLA/TAM

at various temperatures ............................................................................................................... 66

Table 6.1 Mechanical properties of PLA and PLA/PHO blends ..................................................... 76

Table 6.2 Thermal properties of neat, DCP and coagent modified PHO, PLA and PHO/PLA blend

....................................................................................................................................................... 83

Table 6.3 Mechanical properties and heat deflection temperature of neat and coagent modified

PLA, and PLA/PHO blends ............................................................................................................. 90

Page 15: Preparation and characterization of biopolymer compounds

xiv

List of Schemes

Scheme 2.1 Chemical structure of polyhydroxyalkanoate [8] ........................................................ 6

Scheme 2.2 Reaction of diisocyanate with hydroxyl and carboxyl functional groups, adapted

from Lee et al. 2009 [4] ................................................................................................................. 12

Scheme 2.3 Chemical structure of polylactic acid [36] ................................................................. 13

Scheme 2.4 Reaction between PLLA and glycidol adapted from Deenadayalan et al. 2009 [46] 16

Scheme 2.5 General structure of Joncryl® styrene – acrylic multi-functional oligomeric chain

extender, where R1 – R5 are H, CH3, a higher alkyl group, or combinations of them, R6 is an

alkyl group, and X, Y and Z are each between 1 and 20, adapted from Villalobos at al. 2004 [48]

....................................................................................................................................................... 17

Scheme 2.6 Mechanism of PLA reaction with GMA, adapted from Al-Itry et al. 2012 [51] ......... 18

Scheme 2.7 Chemical structure of coagents, adapted from Parent et al. 2008 [56] ................... 19

Page 16: Preparation and characterization of biopolymer compounds

xv

Nomenclature

a Mark-Houwink exponent G* Complex modulus (Pa) G' Elastic or Storage modulus (Pa) G'' Viscous or loss modulus (Pa) K Mark-Houwink constant (dL/g) T Temperature (°C) t Time (s) TC Temperature of crystallization (°C) TM Melting point (°C) Xc Degree of crystallinity (%) Greek Symbols ΔHf Heat of fusion (J/g) [η] Intrinsic viscosity (dL/g) 3η+ Linear viscoelastic envelope in uniaxial extension (Pa.s) η* Complex Viscosity (Pa.s) η0 Zero shear viscosity (Pa.s) ω Angular frequency (rad/s) λ viscosity ratio ηd viscosity of the dispersed phase ηm viscosity of the matrix

Page 17: Preparation and characterization of biopolymer compounds

xvi

Abbreviations

3HV 3-hydroxyvalerate ASTM American society for testing materials CHS cross head speed DBP Di n-butyl phthalate DCP Dicumyl peroxide dn/dc Refractive index DRI Differential refractive index DSC Differential scanning calorimetry GMA Glycidyl methacrylate GMS Glycerol monostearate GPC Gel permeation chromatography GTA Glycerol triacetate GTB Glycerol tributyrate HDI Hexamethylene diisocyanate HDT Heat distortion temperature IV Intrinsic viscosity LALS Low angle light scattering LLDPE Linear low density polyethylene LS Light scattering LVE Linear viscoelastic MCL Medium-chain-length mCPBA m-Chloroperbenzoic acid MK Mark-Houwink MMT Montmorillonite Mn Number-average MSDS Material safety data sheets Mw Weight-average MW Molecular weight MWD Molecular weight distribution NCC Nano-crystalline cellulose NSERC Natural sciences and engineering council PBS Polybutylene succinate PCL Polycaprolactone PDI Poly dispersity index PEG Polyethylene glycol PEGCA Poly (polyethylene glycol-co-citric acid) PEGCA Poly (polyethylene glycol-co-citric acid) PET polyethylene terephthalate PETA Pentaerythritol triacrylate PHA Poly-3-hydroxyalkanoate PHB poly(3-hydroxybutyrate

Page 18: Preparation and characterization of biopolymer compounds

xvii

PHB-HV poly(3-hydroxybutyrate-co-3-hydroxyvalerate PHDD Poly-3-hydroxydodecanoate PHN Poly-3-hydroxynonanoate PHO Poly-3-hydroxyoctanoate PLA Poly (lactic) acid PLLA Poly(L-Lactic Acid) POM Poly(methyleneoxide PP Polypropylene PS Polystyrene PVA Poly vinyl alcohol RALS Right angle light scattering SEC Size exclusion chromatography SEM Scanning electron microscopy SNCC Silylated cellulose nanocrystals TAM Triallyl trimesate TAP Triallyl phosphate TC Crystallization temperature TD Triple detector Tg Glass transition temperature TGA Thermogravimetric analysis THF Tetrahydrofuran TM Melt temperature TMPTA Trimethylol propanetrimethacrylate TPEE Thermoplastic polyester elastomer UC Universal calibration

Page 19: Preparation and characterization of biopolymer compounds

1

Chapter 1 Introduction

The rapid progress in polymer science and technology in the latter half of the twentieth century

has led to the use of synthetic polymers in almost every field, including household items,

industrial applications, electrical appliances, electronic devices, automotive, construction, and

the medical field. Polymer based compounds are found in every aspect of everyday life. The fast

growth of the polymer industry is attributed to the unique properties of polymers including

light-weight, corrosion resistance, design flexibility, and ease of manufacturing. Durability is one

of the advantages of polymers, but it is also their shortcoming. Petroleum based polymers are

generally not biodegradable. Their life span is longer than several hundred years. After the end

of their use, the polymers end up either in landfills or littering the environment. Landfilling is

becoming increasingly prohibitive, due to increasing costs, scarcity of land and other health and

environmental considerations, such as ground water contamination.

Furthermore, conventional plastics are made from petroleum products or natural gases, which

are non-renewable resources. Serious concerns about greenhouse gas emissions [1] and high oil

prices, which accompany the use of conventional polymers, are key driving forces to reduce

their usage. Therefore there is a need for new materials from renewable resources to replace

conventional polymers [2,3]. The quest for alternative materials has put biopolymers at the

forefront [4]. Biopolymers are promising candidates to replace conventional polymers, because

of their biodegradable nature and they can be made from renewable resources as raw material.

They can be categorized as a) bioresourced, b) biodegradable and c) bioresourced and

biodegradable.

Page 20: Preparation and characterization of biopolymer compounds

2

Polylactic acid (PLA) is one of the major bioplastics available on a large commercial scale. It is an

aliphatic polyester made from α hydroxy acids. It is biodegradable and biocompatible and it is

used in many biomedical applications, such as sutures, stents, dialysis media, and drug delivery

devices. PLA is also used in commodity applications ranging from clothing, packaging, bottles,

and office stationary to food containers [5].

Polyhydroxyalkanoate (PHA) is another polymer that has drawn considerable attention. It is

produced by bacteria in a fermentation process, and is a water stable, biodegradable,

biocompatible polymer. Past efforts to commercialize PHA have not been very successful,

because they lack the engineering properties and processability needed to compete with

conventional polymeric materials [6].

1.1 Thesis objectives

Upgrading the properties of biopolymers is an active area of research in academia and industry,

with scientists and technologists striving to match performance of biopolymers with petroleum

based polymers to find new applications. The objective of this thesis is to develop commercially

viable biopolymer formulations containing PHA and PLA. To achieve this goal, medium-chain-

length (MCL) PHA has been identified as a potential impact modifier. Following the detailed

characterization and selection of suitable MCL PHA grades, challenges like the viscosity

mismatch between the polymers and slow crystallization rates will be addressed with the

ultimate goal of developing polymeric material with acceptable engineering properties and

good processability in commercial polymer processing operations.

The approach followed in this thesis is outlined below.

Page 21: Preparation and characterization of biopolymer compounds

3

1.1.1 Characterization of MCL PHA

Even though the synthesis of MCL PHAs has been reviewed extensively, their physical

properties have not been yet fully characterized. Full characterization of the MCL PHAs used in

this study is necessary to further understand their physical properties and processability and to

choose suitable MCL PHA grades for formulations of biopolymers containing these materials.

The first objective of this work is to fully characterize the molecular weight, melt and solid state

properties of various MCL PHA grades, with different chain structures. This will aid with the

choice of materials that will be used in subsequent steps.

1.1.2 Blends of MCL PHA with brittle biopolymers

Biopolymers such as PLA and poly-3-hydroxybutyrate (PHB) are brittle materials and thus

cannot be used in certain applications like packaging, where flexibility of the material is

essential. The second objective is to assess the ability of MCL PHA to impart flexibility into PHB

and PLA. A suitable MCL PHA candidate will be chosen based on properties like melt viscosity,

melting temperature, thermal stability and molecular weight, as described in the first objective.

Various amounts of MCL PHA will be blended with PHB and PLA and the morphology and

mechanical properties of the blends will be evaluated.

1.1.3 Enhancement of melt viscosity of MCL PHA for blending with PHB and PLA

It is expected that the inherently low melt viscosity of MLC PHAs will pose a problem when

trying to blend them with other polymers. The significant difference in the melt viscosity of the

blend components tends to give phase separation, and coarse morphology, resulting in poor

mechanical properties. Therefore, the third objective will be to increase the melt viscosity of

MCL PHAs, using chain extension and cross-linking techniques.

Page 22: Preparation and characterization of biopolymer compounds

4

1.1.4 Improved melt strength and crystallization of PLA and its blends with MCL

PHA by reactive modification

One of the main drawbacks that have hindered widespread implementation of biopolyesters,

including PLA and PHAs, is their low crystallization rates, resulting in poor mechanical

properties and processability. This makes them practically impossible to melt process in a cost-

effective way using conventional techniques like injection molding, compression molding and

extrusion. Poor melt strength restricts PLA’s processability in operations involving high stretch

rates, such as film blowing, thermoforming, and foaming. The last objective of this thesis is to

improve crystallinity and melt strength of biopolymers through reactive processing without

using any nucleating agents.

1.2 Thesis organization

This thesis contains seven chapters. Chapter 1 gives an introduction of biopolymers specifically

PHAs and PLA, their attributes and limitations. The chapter also defines the scope of the

proposed research. Chapter 2 summarizes the literature describing the various approaches that

have been followed to address the limitations of PLA and PHAs. Relevant work is examined

critically. Chapter 3 discusses the characterization of a series of MCL PHAs including true

molecular weight, melt viscosity, thermal stability, glass transition temperature (Tg), melting

temperature and crystallinity. Chapter 4 describes impact modification of brittle PHB, using a

MCL PHA (i.e. polyhydroxyoctanoate (PHO)). Chapter 5 describes the preparation of branched

PLA with improved strain hardening and crystallinity by reactive modification in the melt state,

using a peroxide and a multi-functional coagent. The performance of coagent-modified PLA is

Page 23: Preparation and characterization of biopolymer compounds

5

compared with PLA modified with a multi-functional epoxide styrene-acrylic oligomeric chain

extender, containing glycidyl methacrylate (GMA) functions. Chapter 6 describes impact

modification of PLA, and further utilizes the reactive modification approach described in

chapter 5, to prepare reactively modified PHO and PLA/PHO blends to achieve a good balance

of mechanical and melt-state properties. Chapter 7 summarizes the overall outcomes and

achievements of the thesis with recommendations for future work.

Page 24: Preparation and characterization of biopolymer compounds

6

Chapter 2 Literature Review

2.1 Polyhydroxyalkanoates

Polyhydroxyalkanoates (PHAs) are linear polyesters that were first discovered in late 1920s, by

Lemoigne who produced them using the Bacillus megaterium bacteria [7]. The chemical

structure of PHA is shown in scheme 2.1 [8].

Scheme 2.1 Chemical structure of polyhydroxyalkanoate [8]

PHA can be divided into three categories based on number of carbons in side chain (the R group

in Scheme 2.1): 1) Short-chain-length (SCL) PHA, 2) medium-chain-length (MCL) PHA and 3)

hybrid (mix of SCL PHA and MCL PHA). SCL PHA contains 0-2 carbons whereas MCL PHA

contains 3- 11 carbons in their backbone [9]. They can be found as homopolymers or co-

polymers. PHAs can have thermoplastic or elastomeric properties, depending upon their

composition. SCL PHAs behave like typical brittle thermoplastics, while MCL PHAs are

elastomeric. Melting points of PHAs lie between 40oC - 180oC. PHAs can be processed by

conventional polymer processing techniques.

Poly-3-hydroxybutyrate (PHB), a SCL PHA, is the most widely studied PHA. It is a brittle polymer

and cannot be used without impact modification. The high crystallinity and low crystallization

rate of PHB leads to embrittlement and an ageing effect of the polymer [10]. PHB is not

Page 25: Preparation and characterization of biopolymer compounds

7

thermally stable [11], resulting in narrow processing windows [12]. The brittleness of PHB can

be counteracted by adding impact modifiers or by using copolymers of the same polymer

family, for example poly (3-hydroxybutyrate-co-3-hydroxyvalerate), P (3HB-co-3HV), which is

more ductile and flexible than PHB. Increase in the 3-hydroxyvalerate (3HV) content improves

flexibility with a decrease in melting temperature, tensile strength, modulus of elasticity and

crystallinity. P (3HB-co-3HV) was commercialized in the 1980s by Imperial Chemical Industries

Inc. (ICI) in the UK under the trade name Biopol®.

Medium-chain-length (MCL) PHA

SCL PHAs can be produced using a wide variety of bacteria under nutrient deprivation as an

intracellular energy reserve [13]. On the other hand, very few microorganisms can be used to

produce MCL PHA [14]. Accumulation of MCL PHA is restricted to Pseudomonas rRNA homology

group I, like Pseudomonas aeruginosa, P. chlororaphis, P. putida, P. syringae and some P.

fluorescens [14,15]. P. putida strains (formerly P. oleovorans GPo1) are used widely to produce

MCL PHA. They have the ability to use alkanes, such as octane for synthesis of MCL PHA due to

their octane (OCT) plasmid [13]. Alkanoates such as octanoate are a common carbon source to

make MCL PHAs [16].

The properties of MCL PHAs depend on the polymer composition. The melting point of

polyhydroxyoctanoate (PHO), a MCL PHA, is 61 °C with crystallinity of 30 % [17]. The glass

transition temperature (Tg) is -35 °C. It is elastomeric in nature, having an extension at break

value between 300-450 % and tensile strength between 6 and 10 MPa. The density of PHO is 1

g.cm-3 [17]. Based on its properties listed above, MCL PHA is a potential bio-sourced,

biodegradable impact modifier for brittle biopolymers.

Page 26: Preparation and characterization of biopolymer compounds

8

Additives and chemical modification techniques that are commonly used to enhance the

performance of PHAs and broaden its processing window are outlined below.

2.1.1 Impact modification of PHB

Blending brittle polymers with a ductile polymer, which forms a secondary phase, is the easiest

way to impart flexibility in polymers. Polymer blends can either be miscible or immiscible.

Immiscible blends that are compatible yield properties better than the parent polymers

(synergistic effect) [18]. In the case of incompatible blends, compatibilizers can be used to

enhance the blend properties. Some of the PHA-based blends that have been investigated are

summarized below.

Parulekar et al. [19] used natural rubber to improve the ductility of PHB. They found that PHB-

natural rubber blends were not compatible and there was a substantial viscosity mismatch

between the two polymers. They used maleated polybutadiene with high graft content and low

molecular weight to compatibilize the system. They concluded that epoxidized (25%

epoxidation) natural rubber yielded better ductility compared to the non-functionalized natural

rubber. A maleated polybutadiene- PHB – natural rubber (10:60:30) system showed 440%

improvement in impact strength of PHB.

Block copolymers of polycaprolactone (PCL) and PHB have been used to compatibilize

immiscible PCL/PHB blends. Optimization of the composition of the main blend components

and compatibilizers is crucial to get substantial improvements in properties. 25 wt.% PCL and 5

wt.% of PCL-PHB block copolymers had only an elongation of 29 %, which increased

dramatically to 855 % when there was 45 wt.% of PCL and 10 wt.% of PCL-PHB block copolymer

[21]. Furthermore a 50:50 blend of PHB with a block copolymer of atactic poly ((R, S) -3-

Page 27: Preparation and characterization of biopolymer compounds

9

hydroxybutyrate) and poly (ethylene glycol) resulted in increased elongation at break from 5 %

to 90% with decrease in modulus and tensile strength [22]. An immiscible blend of PHB and

acrylonitrile-g-(ethylene-co-propylene-co-diene)-g-styrene has also been studied and is

composed of four phases of poly (ethylene-co-propylene-co-diene), poly (styrene-co-

acrylonitrile), amorphous PHB and crystalline PHB. The blend exhibits 190% improvement in

impact resistance of pristine PHB [23].

2.1.2 Plasticization

The ductility of PHB can be improved by adding low molecular weight plasticizers. Generally

plasticizers decrease the intermolecular forces between polymer chains, thus increasing chain

mobility. Plasticizers for PHB include 1) high boiling esters of polybasic acids such as phthalates,

isophthalates, citrates, fumarates, glutamate, phosphates or phosphites 2) high boiling esters

and part esters of polyhydric alcohols mainly glycols, polyglycols and glycerol 3) aromatic

sulphonamides [24] and 4) a few high molecular weight polymers.

Ceccorulli et al. studied the effect of plasticization on the Tg of PHB [25]. They used a

biodegradable plasticizer, di n-butyl phthalate (DBP) to improve the ductility of the PHB. 30

wt.% of DBP was sufficient to lower the Tg of PHB from 6 °C to -40 °C. The data is in agreement

with Riande et al. [26] giving evidence of the existence of two concomitant phenomena, Tg

depression of the polymer due to the plasticizing effect with up to 40 wt.% of plasticizer, and a

small increase of the Tg of the plasticizer, due to hindrance in its mobility due to the dissolved

polymer molecule, above 40 wt.% of the plasticizer. Plasticization did not affect the ability of

PHB to crystallize. Increase in the plasticizer content decreases the crystallization temperature

Page 28: Preparation and characterization of biopolymer compounds

10

as a result of reduced Tg that provides enough mobility to macromolecules for rearrangement

and crystallization.

Other plasticizers include glycerol triacetate (GTA), glycerol tributyrate (GTB), and glycerol

monostearate (GMS) [27]. Some high molecular weight polymers can also act as plasticizers. For

example, addition of PEO, which has a Tg of -59 oC lowers the Tg and crystallization temperature

of PHB [20].

2.1.3 Nucleation

One of the main drawbacks that have hindered widespread implementation of biopolyesters,

including PHAs, is their low crystallization rates, which result in poor and inconsistent

mechanical properties and processability. The time that these materials take to crystallize from

the melt can be a few hours, which is too long compared to most conventional polymers. This

makes them practically impossible to melt process in a cost-effective way using conventional

processing techniques. Improvements in the crystallization behavior are necessary to obtain

biopolymer-based formulations with good processability in polymer processing operations,

such as extrusion, injection molding, and film processing.

Nucleating agents are commonly used to overcome slow crystallization rates and to increase

the overall crystallinity, while maintaining small crystal sizes to achieve improved mechanical

properties. Some of the common nucleating agents used in industry include talc, mica, calcium

carbonate, chalk, and boron nitride. Environmentally friendly nucleating agents include

saccharin, and phthalimide but they are not as effective as the conventional nucleating agents,

such as boron nitride. They are both soluble in the melt and crystallize when solubility exceeds

the relevant threshold [28].

Page 29: Preparation and characterization of biopolymer compounds

11

High molecular weight polymers like poly vinyl alcohol (PVA) can also be used to nucleate PHB.

PVA is a biodegradable, biocompatible and water soluble polymer with high crystallinity. A lot

of research has been done on blends of PHB and PVA, which form a miscible blend with lower

crystallinity. Alata et al. [29] showed that PVA particles can act as nucleating agents for PHB,

and enhance the rate of crystallization of PHB. PVA has a melting point of ~ 225 °C, so blends

processed at 190 °C ensure melting of PHB while PVA particles are still in the solid state. The

nucleating performance of PVA is equivalent to that of talc and provides a complete bio and

environment friendly material. Nano clays have also been used as nucleating agents for PHB.

Addition of montmorillonite (MMT) clay increased the crystallization temperature by 29 °C [30].

In search of bio-sourced alternatives, which are considered more sustainable, cellulose, which is

a naturally available crystalline material derived from wood (wood contains 40-50% cellulose

[31]), has been suggested as a potential candidate. Cellulose has been proven to act as a

nucleating agent for polymers like polypropylene. It reduces the size of the spherulites, while

increasing the overall crystallinity and inducing trans-crystallinity [32,33]. Most of the work

reported in this area has been done on microcrystalline cellulose. With recent advances in

nanotechnology, it has been shown that high-aspect ratio nano-sized fillers are advantageous

compared to their micron sized counterparts. Desired property improvements can be obtained

at a fraction of the loading, while loss of ductility is minimized. Nano-crystalline cellulose (NCC)

is a completely bioderived and biodegradable material, derived from cellulose through acid

hydrolysis. NCC has been the subject of many recent research initiatives, including partnerships

with the Canadian government [34]. In addition to being a promising candidate as a nucleating

agent for PHAs, it can also offer substantial reinforcement given its high strength [35].

Page 30: Preparation and characterization of biopolymer compounds

12

2.1.4 Chain extension

Chain extension is used widely to increase the molecular weight of polyesters. Diisocyanate

forms urethane and amide linkages through a reaction with hydroxyl and carboxyl functional

group, resulting in significant increase in molecular weight. Hexamethylene diisocyanate (HDI)

has been used as a chain extender for PHO [4]. Scheme 2.2 illustrates the reaction of hydroxyl

and carboxyl functional groups of polyester with isocyanate forming urethane, amide and

allophanate linkages. This reaction increased the weight average molecular weight of PHO by

275 % and its number average molecular weight by 314 % [4].

Scheme 2.2 Reaction of isocyanate with hydroxyl and carboxyl functional groups, adapted from

Lee et al. 2009 [4]

2.2 Polylactic acid (PLA)

Poly (lactic acid) (PLA) (scheme 2.3) is an aliphatic polyester derived from renewable resources.

Page 31: Preparation and characterization of biopolymer compounds

13

Scheme 2.3 Chemical structure of polylactic acid [36]

The properties of PLA depend significantly upon its molecular weight and the stereochemical

makeup of the backbone, which is controlled by polymerization with D-lactide, L-lactide, or D,L-

lactide, to form random or block stereocopolymers [37]. Some of the general properties of PLA

are summarized in Table 2.1

Table 2.1 Properties of PLA [37]

PLA

Density (Kg/m3) 1.26

Tensile strength (MPa) 59

Elastic modulus (GPa) 3.8

Elongation at break (%) 4-7

Notched izod (J/m) 26

Heat deflection temperature (°C) 55

Some of the limitations of PLA are quite similar to PHB, for example brittleness and low

crystallization rate. PLA has a very narrow processing window, because of the lack of melt

strength and its slow crystallization rates. Its poor engineering properties, including impact

strength and heat resistance have mainly confined its applications to food packaging [38], as

Page 32: Preparation and characterization of biopolymer compounds

14

well as biomedical applications, such as drug delivery, where biocompatibility and

biodegradability are desired. Extensive research has been done on PLA to address these

problems. Various approaches to address the challenges of PLA are discussed below.

2.2.1 Impact modification

Blending with ductile polymers or elastomers is commonly used to counteract the brittleness of

PLA. Thermoplastic polyester elastomer (TPEE) has been used as an impact modifier for PLA.

Forming an immiscible blend with two phase morphology 4, 4-Methylenebis(phenylisocyanate)

acts as a compatibilizer increasing the interfacial adhesion between PLA and TPEE to give 340 %

increase in elongation at break while also maintaining modulus and tensile strength [39].

MCL PHA has been tested as an impact modifier for PLA. Immiscible blends of MCL PHA and

PLA, produced by solution mixing, exhibited substantial improvement in impact strength and

decreased tensile strength compared to the neat PLA [9]. Epoxy functionalized MCL PHA gave

further improvement in mechanical properties. The epoxy groups react with the hydroxyl group

of PLA, thus increasing the interfacial interaction and improving the blend morphology. Chain

extension reduced the viscosity mismatch between PLA and PHO, so that eventually the melt

viscosity of PHO and PLA blends was higher than that of blends without chain extension [4].

2.2.2 Plasticization

Polyethylene glycol (PEG) is often used to plasticize PLA. The system has improved elongation at

break with limited impact strength. The PEG based polyester, poly (polyethylene glycol-co-citric

acid) (PEGCA) forms a partially miscible blend with PLA. Addition of PEGCA diminishes the Tg of

PLA. At 15 wt.% it gives 242 % elongation at break with impact strength as high as 103 J/m [40].

Page 33: Preparation and characterization of biopolymer compounds

15

Adipates like (bis (2-ethylhexyl) adipate and glyceryl triacetate) and polymeric adipates exhibit

excellent plasticization of PLA composites. They decrease Tg of PLA by 20 °C. 10 wt. % of

plasticizer gives four times higher impact properties for PLA composite containing 40 wt. % of

stable β -anhydrite. Adipates also help improve the dispersion of fillers and results in better

tensile strength of composites [41].

2.2.3 Nucleation

Nucleating agents are used to address the slow crystallization rate of PLA. They increase

nucleating density and thus increase crystallization rate. Talc exhibits effective nucleation in

PLA, whereas other nucleating agents like calcium lactate and sodium stearate had little or no

nucleating effect on PLA [42]. Talc contents of 1 wt.% significantly accelerate the crystallization

process of the PLA matrix. The maximum crystallization rate was observed at an annealing

temperature of 100 °C [43]. NCC has also been assessed in PLA. It has been shown that silylated

cellulose nanocrystals (SNCC) exhibits enhanced nucleating efficiency. Addition of 1 wt.% SNCC

resulted in substantial increase in crystallization rate. Increased crystallinity resulted in

improvements of tensile strength and modulus [44].

2.2.4 Conditioning

Annealing is another approach to toughen PLA. The annealing temperature and annealing time

have a significant effect on crystallization. As annealing time and temperature increase, the

impact strength increases due to smaller spherulite size and a larger amount of the metastable

phase. Quenching and subsequently annealing at an appropriate temperature (~90 °C) is crucial

for PLA toughening [44].

Page 34: Preparation and characterization of biopolymer compounds

16

2.2.5 Chain extension

As most polyesters, PLA has poor melt strength which restricts its use in processes involving

extensional flow, such as film processing, thermoforming, and foaming. The processing window

of PLA can be broadened by chain extension. Chain extenders such as isocyanate, glycidol,

peroxides, and epoxy based styrene-acrylic oligomers have been used to increase the molecular

weight of PLA, to improve its thermal stability and to increase its melt strength by introducing

strain hardening [45], as explained in detail below.

2.2.6 Epoxy based chain extension

2.2.6.1 Glycidol

Deenadayalan et al. [46] achieved chain extension of PLA by reactive extrusion in the presence

of glycidol. Chain extension was initiated by reaction of the carboxyl and the hydroxyl of PLA

end groups with glycidol. The carboxylic end group of Poly(L-Lactic Acid) (PLLA) reacts with the

primary hydroxyl end groups of glycidol and forms end-capped linear PLLA (scheme 2.4).

Scheme 2.4 Reaction between PLLA and glycidol adapted from Deenadayalan et al. 2009 [46]

Page 35: Preparation and characterization of biopolymer compounds

17

A hydroxyl end group from another PLLA reacts with glycidol and initiates chain extension with

formation of a pendant hydroxyl group that can react further with another end-capped PLLA to

form a branch structure. Using this reaction, the resulting increase in molecular weight was

more pronounced when low molecular weight as compared to high molecular weight PLA was

used; this was attributed to the higher concentration of end groups in low molecular weight

PLA, resulting in improved chain extension. The modified PLA showed higher Tg and melt

temperature.

2.2.6.2 Multi-functional epoxy based chain extenders

Joncryl®, a multi-functional epoxide styrene-acrylic oligomeric chain extender, containing

glycidyl methacrylate (GMA) functions (scheme 2.5) has been used successfully as a chain

extender for polyesters, such as PLA [47]. It has following physical characteristics: Tg – 55 °C,

epoxy equivalent weight – 285 g/mol.

Scheme 2.5 General structure of Joncryl® styrene – acrylic multi-functional oligomeric chain

extender, where R1 – R5 are H, CH3, a higher alkyl group, or combinations of them, R6 is an

alkyl group, and X, Y and Z are each between 1 and 20, adapted from Villalobos et al.2004 [48]

Page 36: Preparation and characterization of biopolymer compounds

18

GMA contains epoxy groups, which participate in reactions with hydroxyl and carboxyl groups

[49]. In polyesters, glycidyl esterification of carboxylic acid leads to hydroxyl end group

etherification. Reaction of hydroxyl end group etherification competes with etherification of

secondary hydroxyl group and main chain trans-esterification. The epoxy ring opening reaction

results in covalent bonds via hydroxyl side group formation [50]. The reaction mechanism

proposed by Al-Itry is depicted in scheme 2.6 [51].

Scheme 2.6 Mechanism of PLA reaction with GMA, adapted from Al-Itry et al. 2012 [51]

2.2.7 Peroxide-mediated cross-linking

Peroxide curing is one of the oldest technologies to introduce branching and/or cross-linking in

polymers. It is used widely in elastomers and polyolefin technology. A general scheme of

peroxide cross-linking is shown in section 2.2.8. The peroxide undergoes hemolytic

decomposition upon heating to generate free radicals. Free radicals abstract hydrogen from the

polymer chain creating polymer radicals and decomposition products. The polymer radicals can

Degradation

GMA

GMA

Acid End GroupVinylic terminated ester

PLA

Page 37: Preparation and characterization of biopolymer compounds

19

react with each other to form a C-C crosslinks. A competitive degradation reaction can also take

place through β-chain session.

In PLA, peroxide-mediated chain extension has been achieved by reactive extrusion using

lauroyl peroxide [52] dicumyl peroxide [53] and di-tertiary alkyl peroxide [54] with limited

success. The choice of peroxide typically depends on the peroxide half-life time and the

polymer processing temperature.

2.2.8 Multifunctional coagents

Multifunctional coagents are used widely to form branched polyolefins [55,56]. Parent et al.

[57] studied triallyl trimesate (TAM), trimethylolpropane triacrylate (TMPTA) and triallyl

phosphate (TAP) coagents (Scheme 2.7) in polypropylene. Depending on the type of coagent,

different molecular weight and branching distributions may be obtained.

Scheme 2.7 Chemical structure of coagents, adapted from Parent et al. 2008 [56]

A typical peroxide assisted coagent modification scheme is illustrated below.

Page 38: Preparation and characterization of biopolymer compounds

20

Initiation

ROOR 2 RO•

RO• + P-H ROH – P•

Propagation

P• + M. → P-M•

P-M• + P-H → P-MH + P•

Chain transfer

P• + P-H P-H + P•

Fragmentation -

P• P= + P•

Termination

P• + P• → P-P or P-H + P=

More specifically, the TAM activation reaction is discussed by Parent et al. [56].

Page 39: Preparation and characterization of biopolymer compounds

21

Chapter 3 Determination of Mark-Houwink parameters and absolute

molecular weight of medium-chain-length poly(3-

hydroxyalkanoates)*

3.1 Introduction

Current global production of commodity plastics consumes up to 270 million tons of petroleum

annually [58]. Despite recycling efforts, as much as 50% of the 100 million tons of plastic

produced annually may end up in landfill sites. To address environmental and resource issues,

"bio-based plastics” that are both biodegradable and made from renewable resources are being

developed for a variety of applications. New applications and innovations are anticipated in the

automotive, electrical/electronic, medical and packaging industries. Bioplastics include bio-

polyethylene, polylactic acids (PLAs), polyhydroxyalkanoates (PHAs), and starch-based

materials. Forecast analysts including Technavio and Helmut Kaiser Consultancy predict that the

bioplastics market will experience significant growth (33.9-41% compound annual growth rate

(CAGR)) from 2010 to 2015 [59].

As mentioned in section 2.1, the physical properties and solubilities of PHA are greatly affected

by the length of the side group R (Scheme 2.1). Poly(3-hydroxybutyrate) (PHB), the most

common short-chain-length (SCL) PHA, is stiff and brittle, whereas medium-chain-length (MCL)

PHAs are more elastic and flexible. MCL PHAs are biodegradable materials with low crystallinity,

low glass transition temperature (Tg), high elongation, and low tensile strength [60], making

them good candidates for applications where elastomeric properties are needed, such as for

*A version of this chapter has been published. Nerkar, M., Ramsay, J.A., Ramsay, B.A., Kontopoulou, M., Hutchinson, R.A. Journal of Polymers and the Environment 2013 (21): 24-29

Page 40: Preparation and characterization of biopolymer compounds

22

impact modification of brittle biopolymers [47,61]. MCL PHAs are the only thermoplastics

elastomers that have 100 % renewable content and are biodegradable.

Given the influence on many engineering properties such as strength, stiffness, toughness,

elasticity, viscosity and thermal transitions, it is important to have access to reliable molecular

weight data for PHAs, to assess the effect of synthesis conditions on their properties and

evaluate the potential of these biopolymers in engineering applications.

Of the various PHA materials, PHB is by far the most investigated and characterized polymer. As

PHB is not soluble in tetrahydrofuran (THF), molecular weight (MW) characterization is done in

more aggressive solvents such as chloroform, 2,2,2-trifluoroethanol, and ethylene dichloride.

Akita et al. [62] and Marchessault et al. [63] employed light scattering and osmometry to obtain

absolute molecular weight (MW) data and to determine the Mark-Houwink (MH) calibration

constants required to estimate PHB molecular weight distributions (MWD) from size exclusion

chromatography (SEC) analysis coupled with a single detector calibrated using polystyrene (PS)

standards. Ubbelohde type capillary viscometry and rotational viscometry have also been used

to determine intrinsic viscosity data [62]. Miyaki et al. [64] and Cornibert et al.[65] covered a

broader molecular weight range using osmometry. These efforts allow the estimation of

absolute MW values from single-detector SEC analysis: a value of 1.0105 Da as measured by PS

calibration is transformed to 0.6105 Da for PHB in chloroform using the MH constants

published by Akita et al., a shift of 35%. PHB molecular weight is typically in the order of 1106

Da in native granules but may decrease if exposed to depolymerases, base, oxidizing agents

such as hypochlorite [66] or chlorinated solvents [67].

Page 41: Preparation and characterization of biopolymer compounds

23

However, only limited MW data is available in the literature for MCL PHAs [60] and MCL SCL

PHA materials [68], as measured by SEC analysis in THF eluent, and MW values are typically

reported relative to PS calibrations. Absolute values of molecular weight for MCL PHAs are not

reported in the literature to date, as the necessary SEC calibrations had not been performed.

This information is required to better relate polymer structure to the processing properties and

the degradation kinetics of these emerging biomaterials, such as those reported by Daly et

al.[69] for poly (3-hydroxybutyrate-co- 3-hydroxyhexanoate).

In this chapter, absolute MW averages have been determined for four different MCL PHA

copolymers from MWDs measured using multi-detector SEC, and MH parameters were

estimated for the first time for these kinds of polymers in THF. Furthermore the melt viscosity is

reported as a function of molecular weight.

3.2 Experimental

3.2.1 Materials

The four samples were copolymers with (3-hydroxyoctanoate) (PHO), (3-hydroxynonanoate)

(PHN) and (3-hydroxydodecanoate) (PHDD) moieties produced from renewable starting

materials including sugar and vegetable oil. PHO contained 98 mol % 3-hydroxyoctanoate and 2

mol % 3-hydroxyhexanoate. PHN 90 and PHN 70 contained 3-hydroxynonanoate and 3-

hydroxyheptanoate at mole ratios of 90/10 and 70/30, respectively. PHDD contained 40 mol %

3-hydroxydodecanoate 39 mol % 3-hydroxydecanoate, 19 mol % 3-hydroxyoctanoate and 2 mol

% 3-hydroxyhexanoate.

Page 42: Preparation and characterization of biopolymer compounds

24

PHN 90 and PHO were produced in chemostat culture with addition of acrylic acid to inhibit β-

oxidation [70]. PHN 70 and PHDD were produced in fed-batch culture [71]. The polymer

samples were extracted from washed lyophilized biomass by Soxhlet extraction in acetone

followed by precipitation in cold methanol [72], except for PHDD which was extracted in

chloroform [67]. Cellular PHA content and composition were determined by gas

chromatography as described by Sun et al. [71].

3.2.2 Methods

Samples were prepared for SEC analysis by dissolving 10 mg of polymer in 1 mL of distilled THF

overnight to ensure complete dissolution, then passed through a 0.2 µm nylon filter. Polymer

molecular weight distributions were measured using two different instruments to check

consistency and reliability of the data. The first, a Viscotek 270max separation module with

triple detection by differential refractive index (DRI), viscosity (IV) and light scattering (low

angle LALS and right angle RALS), was maintained at 40 °C and contained two porous

PolyAnalytik columns in series with an exclusion molecular weight limit of 20106 Da. Distilled

THF was used as the eluent at a flow rate of 1 mL/min. The MWDs were calculated using two

methodologies. First, the results from the triple detector train and Viscotek Omnisec

software were used to determine polymer MWDs and MW averages using the values of the

refractive index (dn/dc) determined offline, using a Wyatt Optilab DSP refractometer as

described below. The triple detection mode also yielded estimates for the polymer Mark-

Houwink (MH) parameters, determined directly from the curve generated by the output from

the IV and LS detectors. These MH parameters provided the means for a second analysis of the

output data using the principle of universal calibration and the output from the DRI detector,

Page 43: Preparation and characterization of biopolymer compounds

25

with a calibration curve constructed with narrow molecular weight polystyrene standards

ranging from 6910 to 3.3106 Da.

Samples were also characterized on a second SEC instrument consisting of a Waters 2960

separation module coupled with a Waters 410 differential refractometer (DRI) and a Wyatt

Instruments Dawn EOS 690 nm laser photometer multiangle light scattering (LS) detector. THF

was used as eluent at a flow rate of 1 mL/min through the four Styragel columns (HR 0.5, 1, 3,

4), maintained at 35 °C. The DRI detector was calibrated by 10 narrow polydispersity

polystyrene standards in a broad MW range (870-3.55105 Da), and the LS detector was

calibrated by toluene, as recommended by the manufacturer. MWDs from this instrument were

again calculated using two methods, using the output from the DRI detector and the MH

parameters determined with the Viscotek setup, and by processing the data from the LS

detector using the Wyatt Astra software and the refractive index (dn/dc) of the polymer in THF,

according to standard procedures. The refractive index values were measured by a Wyatt

Optilab DSP refractometer at 35 °C and 690 nm calibrated with sodium chloride. Five samples of

3–18 mgmL–1 were prepared in THF for each polymer and injected sequentially to construct a

curve with slope dn/dc [73]. The values for PHN 90 and PHO were 0.06010.0002 mL/g and

0.06030.0003 mL/g, respectively.

DSC experiments were performed using a Q100 DSC from TA Instruments, under dry nitrogen.

Since MCL PHAs crystallize slowly, the samples were preconditioned to eliminate their thermal

history as follows: the polymer was heated at 100 °C for 10 min in a convection oven, and then

kept at room temperature for two weeks before characterization. Samples weighing 10-12 mg

were sealed in aluminum hermetic pans, equilibrated at -70 °C and kept isothermally for 5 min.

Page 44: Preparation and characterization of biopolymer compounds

26

Afterwards they were heated to 100 °C at a rate of 5 °C/min and held isothermally for 3 min

before cooling to -70 °C at a rate of 5 °C/min. As MCL PHAs did not crystallize during second

heating cycle data from first heating cycle was used for differentiation. The samples were finally

reheated to 100 °C at a rate of 5 °C/min. The % crystallinity of the polymers, Xc, was estimated

using equation (3.1).

100H

HX

100

mc

(3.1)

where, ΔHm is the enthalpy of fusion and ΔH100 is the theoretical fusion enthalpy of a 100%

crystalline polymer, which is 146 J/g [74].

Rheological characterization was carried out in the constant rate mode under nitrogen blanket

using a ViscoTech rheometer by Reologica, equipped with a cone and plate fixture having 25

mm diameter at 120 °C. All MCL PHAs had Newtonian behavior.

3.3 Results and Discussion

3.3.1 Molecular weight determination

Two different instruments, a triple detector Viscotec 270 max and a dual detector

Waters/Wyatt, were used to characterize the polymer samples, as detailed in the Experimental

Section. MWDs and molecular weight averages were calculated from each instrument using

both multidetector and single detector (DRI) with universal calibration) output.

Log-log plots of polymer intrinsic viscosity ([] in dL/g) vs. MW were constructed using the

triple detector output from the Viscotec SEC; experimental results for the three PHO replicates

are shown in Figure 3.1a. These curves were calculated using the dn/dc value of 0.0602 mLg–1

Page 45: Preparation and characterization of biopolymer compounds

27

determined to be independent of the MCL PHA composition in THF at 35 °C. Linear regression

was used to estimate the MH calibration parameters and for all four samples, as summarized in

Table 3.1. Averages and standard deviations of three measurements are reported. The four sets

of calibration parameters represent a very similar [] vs MW behavior for the copolymer

samples (Figure 3.1b). Thus, the values were averaged to provide an estimate of a “universal”

set of MH parameters that may be applied to all MCL PHAs, independent of composition. The

last column in Table 3.1 reports MW values calculated for a PS equivalent MW of 105 Da using

the four individual pairs of MH parameters as well as that calculated according to the

“universal” MH parameters. The difference between the various estimates is less than 15%,

which is the typical error reported for MW measurements by SEC. Moreover, the calculated

values are within 10% of 105 Da, the PS equivalent MW.

All of the copolymers were characterized at least three times on three different days to check

reproducibility of the data. The MWD data of PHO (three replicates run on two instruments

analyzed using multidetector and universal calibration) are shown in Figure 3.2. The peak

positions of the three replicates are tightly grouped at 105 Da, independent of the SEC setup or

analysis method chosen. This agreement, also observed for the other samples, demonstrates

the validity of the polymer refractive index (dn/dc) and MH calibration parameters determined

in this work for MCL PHA materials.

Before drawing generalized conclusions, it is useful to examine the number-average (Mn) and

weight-average (Mw) molecular weight values as well as the dispersity index (PDI= Mw/Mn) for

the four MCL PHA samples. Only the output from the Viscotek instrument was used, as the

column set used in the Waters/Wyatt instrument is designed for analysis of lower MW samples

Page 46: Preparation and characterization of biopolymer compounds

28

and the calibration for the DRI detector only extends to 3.5105 Da. Thus, the polymer MWDs

are cut off at higher molecular weights, most clearly seen in Figure 3.2a with the DRI detector.

Figure 3.1 a) Experimental Mark-Houwink data for PHO (three replicates), b) best fit Mark-

Houwink relationships for all copolymer samples.

(The Wyatt ASTRA software used for processing of the LS detector provides an estimate of the

complete MWD; however, separation of the higher MW polymer will be incomplete due to the

column set used.) MW averages and standard deviations are reported in Table 3.2 as estimated

a)

b)

Page 47: Preparation and characterization of biopolymer compounds

29

by the full Viscotek triple detector (TD) analysis, as well as using only the output from the DRI

detector coupled with universal calibration (DRI/UC). The universal calibration procedure is

done using the sample-specific MH parameters reported in Table 3.1, as well as the averaged

“universal” MH sample-specific MH parameters reported in Table 3.1, as well as the averaged

“universal” MH parameters. Furthermore, the MW averages are also reported according to DRI

analysis with PS calibration.

Table 3.1 Mark Houwink calibration constants from best fit to triple detector analysis of MCL

PHA samples in THF ([Ƞ]=K(MW)a).

a Calculated for PS equivalent MW of 105 Da

As summarized in Table 3.2, the MWDs of the four samples have similar PDI values of 1.8-2.3.

(The values obtained by TD-SEC are lower, as is often observed in MWDs measured by light-

scattering techniques). However, as also seen in the MWDs plotted as Figure 3.3, the absolute

a Log(K/(dL·g-1)) MW (kDa)a

PS [75] 0.716 -3.943 100.0

PHO 0.701 (0.013) -3.77 (0.07) 88.1

PHN 90 0.663 (0.020) -3.62 (0.11) 92.7

PHN 70 0.701 (0.031) -3.86 (0.16) 99.8

PHDD 0.691 (0.013) -3.86 (0.07) 106.0

Universal PHA 0.689 -3.78 96.2

Page 48: Preparation and characterization of biopolymer compounds

30

Mw averages vary from 3.0-3.4104 Da for the PHN-70 sample to 1.4-1.7105 Da for PHO. These

values, controlled by the synthesis conditions for the different copolymers, are within the range

Figure 3.2 Molecular weight distributions for PHO (three replicates) as measured by a)

Waters/Wyatt SEC, light-scattering detector; b) Waters/Wyatt SEC, universal calibration; c)

Viscotek SEC, triple detection; and d) Viscotek –SEC, universal calibration.

a) b)

c) d)

Page 49: Preparation and characterization of biopolymer compounds

31

Table 3.2 Number-average (Mn), weight-average (Mw), and dispersity (PDI) values determined from SEC analysis of MCL PHA

samples.

Values are reported according to polystyrene (PS) calibration, universal calibration (UC) with individual polymer MH parameters

from Table 3.1, UC with “universal” parameters for MCL PHA, and triple detector (TD) analysis. The average of three samples is

reported.

PS Calibration UC, individual MH parameters UC, averaged MH parameters TD Analysis

Mn (kDa) Mw (kDa) PDI Mn (kDa) Mw (kDa) PDI Mn (kDa) Mw (kDa) PDI Mn (kDa) Mw (kDa) PDI

PHO 73.74.5 15016 2.020.11 61.65.7 140.715.3 2.280.22 67.26.3 155.217 2.310.22 98.2±6.7 172±17 1.75±0.10

PHN90 52.71.6 98.32.6 1.87010 44.92.0 92.62.4 2.060.14 47.42.1 95.62.4 2.030.14 49.8±1.8 89.7±7.0 1.80±0.14

PHN70 15.40.7 33.60.5 2.180.11 13.60.8 31.10.4 2.290.14 12.90.7 29.90.4 2.320.14 18.2±3 31.9±2.3 1.77±0.17

PHDD 25.71.4 47.11.4 1.830.05 24.51.6 47.31.5 1.930.07 22.31.4 43.21.3 1.930.07 29.7±1.3 46.8±0.8 1.58±0.04

Page 50: Preparation and characterization of biopolymer compounds

32

Figure 3.3 Molecular weight distributions of four MCL PHA polymers (Viscotek SEC, universal

calibration).

reported in the literature for MCL PHAs [60,68].The difference in Mw values calculated using TD-

SEC and DRI/UC methodologies is the largest for PHO, with the spread for the other three

(lower Mw) samples less than 10%. This very good agreement (Table 3.1) validates the adoption

of a “universal” set of MH parameters for MCL PHA samples analyzed in THF, and is

independent of copolymer composition.

3.3.2 Thermal and rheological characterization

The thermal properties of MCl PHAs are summarized in Table 3.3. All the materials were highly

amorphous, with PHO having the highest melting temperature and lowest crystallinity, whereas

PHDD has highest crystallinity.

Page 51: Preparation and characterization of biopolymer compounds

33

Table 3.3 Thermal properties of MCL PHAs

TM (°C) Tg (°C) Crystallinity (%)

PHO 63 -36 15 PHN 990 60 -46 17 PHN 970 50 -47 16 PHDD 61 -40 21

The bulk rheology of the four samples was also measured. All samples demonstrated

Newtonian behavior. The zero shear viscosity scales with Mw with a slope of 1 at lower

molecular weights, and has a slope of 3.8 which is higher than the slope of 3.4 anticipated for

linear polymers at the higher molecular weights, as shown in Figure 3.4. This may be due to

experimental error. The plot suggests an entanglement molecular weight around 8104 Da

(where the two fitted lines intersect), which is significantly higher than most conventional

polymers, suggesting that these polymers may adopt folded helical conformations, similar to

what has been proposed for PHB [76].

Figure 3.4 Viscosity as a function of molecular weight

1

10

100

1.E+04 1.E+05 1.E+06

η ( P

a S

) a

t 1

20

°C

Mw (Da)

η~Mw

η~Mw3.8

Page 52: Preparation and characterization of biopolymer compounds

34

3.4 Conclusion

This study provides the first determination of absolute MWDs and MW averages of MCL PHA

copolymers. Fortuitously (and unlike PHB in chloroform), the relationship between polymer

MW and intrinsic viscosity is very close to that determined through PS calibration. Thus,

previously reported MW data [67,77] for MCL PHAs relative to polystyrene calibration can be

considered, within experimental error, as absolute values. With these results, it will be possible

to more closely examine the relationship between MCL PHA synthesis conditions and polymer

MWs, and to better assess their processability using viscosity data. The latter will be of

significant benefit in product development.

Page 53: Preparation and characterization of biopolymer compounds

35

Chapter 4 Melt compounded blends of short and medium-chain-length

poly-3-hydroxyalkanoates*

4.1 Introduction

Poly(3-hydroxyalkanoates) (PHAs) are microbially produced, biodegradable polymers derived

from renewable resources [6]. Poly-3-hydroxybutyrate (PHB), the most studied short-chain-

length (SCL) PHA, is a brittle polymer requiring modification to render it suitable for various

engineering applications. Incorporation of 3-hydroxyvalerate (3HV) to form a copolymer of

poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHB-HV) results in improved ductility, impact

properties, and lower tensile strength, moduli, melting temperatures and crystallinity [78].

Blends of PHB with other polymers have been studied extensively [79,80]. Parulekar et al. [19]

achieved 440% improvement in the impact strength of PHB by adding a maleated

polybutadiene to PHB/natural rubber blends. Substantial improvements in the impact

resistance of PHB have been noted in immiscible blends of PHB with acrylonitrile-g-(ethylene-

co-propylene-co-diene)-g-styrene, comprising of four phases: poly (ethylene-co-propylene-co-

diene), poly (styrene-co-acrylonitrile), amorphous PHB and crystalline PHB [23].

Polycaprolactone (PCL)/PHB blends containing a PCL-PHB block copolymer as compatibilizer,

exhibited dramatic increases in the elongation at break [21]. Blends of PHB containing a block

copolymer of atactic poly((R, S)-3- hydroxybutyrate) and poly(ethylene glycol) resulted in

increased elongation at break and lower modulus and tensile strength [22].

PHB/poly(methyleneoxide)(POM) blends exhibited modest improvements in mechanical

*A version of this chapter has been published. Nerkar, M., Ramsay, J.A., Ramsay, B.A., Kontopoulou, M., Journal of Polymers and the Environment, 2014 (22): 236-243

Page 54: Preparation and characterization of biopolymer compounds

36

properties [20]. Blends of PHB with polymers derived from renewable sources, such as

poly(lactic acid) (PLA), starch and chitosan have also been studied [80-82].

Medium-chain-length (MCL) PHAs, such as poly (3-hydroxyoctanoate) (PHO), have low

crystallinity and exhibit elastomeric properties. PHO has been used to impact modify PLA via

solution blending [9]. Epoxy functionalized MCL PHA resulted in further improvements in

mechanical properties. These materials are therefore promising biopolymers as impact

modifiers for PHB [60]. Dufresne et al. [83] noted a transition from elastomeric to brittle

properties upon addition of PHB to PHO using a solvent mixing technique. Martelli et al. [84]

noted 50% improvement in elongation at break of PHB-HV upon addition of MCL PHA using

solution casting. However melt compounding, which is more industrially relevant, has not been

investigated, in part due to the technical difficulties in producing sufficient quantities of MCL

PHAs.

The objective of this chapter is to prepare and characterize melt compounded blends of MCL

PHAs with PHB. In the previous chapter the properties of four different MCL PHAs with

predominantly 3-hydroxyoctanoate (PHO), 3-hydroxynonanoate (PHN) or 3-

hydroxydodecanoate (PHDD) content were reported. PHO had the highest melt viscosity and

molecular weight. It also has low crystallinity and glass transition temperature (Tg). Based on

these attributes PHO was chosen as the preferred candidate for impact modification in the

present chapter. The morphology, mechanical, thermal and rheological properties of PHB/PHO

blends are reported. Furthermore peroxide-initiated cross-linking of PHO is used to counteract

the viscosity mismatch between the two components.

Page 55: Preparation and characterization of biopolymer compounds

37

4.2 Experimental

4.2.1 Materials

An unmodified, additive free PHB (grade T19), with a weight-average molecular weight of

1,416,000 Da with a dispersity of 7, was supplied by BIOMER, Krailling, Germany in the form of

powder. PHO, containing 98 mol % 3-hydroxyoctanoate and 2 mol % 3-hydroxyhexanoate, was

produced and characterized as described previously in chapter 3 [85]. Its weight average

molecular weight, determined by triple-detector size exclusion chromatography (SEC), was

172,000 Da with a dispersity of 1.75. Lauroyl peroxide (L-231) and ethyl ether anhydrous were

obtained from Elf Atochem and Sigma Aldrich respectively and were used as received.

4.2.2 Compounding

PHB and PHO were dried in a vacuum oven at 100 °C and at room temperature respectively, to

remove moisture. Blends containing 0-30 wt.% PHO were compounded in a DSM

microcompounder at 190 °C for 3 min at a screw speed of 100 rpm. The compounder was

operated under nitrogen blanket to limit polymer degradation. After compounding, the strands

were quenched in cold water before chopping into pellets.

4.2.3 PHO cross-linking

Weighed amounts of lauroyl peroxide (0.06-0.5 wt.%) were dissolved in anhydrous ethyl ether

and coated onto PHO pellets in a glass Petri dish. The coated pellets were placed in a vacuum

oven overnight at room temperature to remove the solvent. Cross-linking was conducted in a

Carver press at 155oC for 10 min to ensure complete reaction (the half-life time of the lauroyl

peroxide at 155oC is 0.8 min). Cross-linked PHO was chopped into small pieces and dry mixed

with PHB at appropriate ratios before feeding into compounder. For rheological

Page 56: Preparation and characterization of biopolymer compounds

38

characterization, the cross-linked compression molded sheets were cut into 25 mm diameter

disc and used for testing as described below.

The gel content of the peroxide-cross-linked MCL PHA was measured by dissolving the material

in boiling tetrahydrofuran (THF) for 7 h. The polymer was sealed in stainless steel wire mesh

(120 mesh) according to ASTM D 2765. The material was left for 1 h under the fumehood and

subsequently dried overnight in a vacuum oven at room temperature. The % gel content was

calculated using equation (4.1).

100sample of weight Initial

sample of weight Finalcontent Gel (4.1)

4.2.4 Blend Characterization

4.2.4.1 Differential scanning calorimetry (DSC)

DSC experiments were performed using a Q100 DSC from TA Instruments, under dry nitrogen.

Since MCL PHAs crystallize slowly, the samples were preconditioned to eliminate their thermal

history as follows: the polymer was heated at 100 °C for 10 min in a convection oven, and then

kept at room temperature for two weeks before characterization. Samples weighing 10-12 mg

were sealed in aluminum hermetic pans, equilibrated at -70 °C and kept isothermally for 5 min.

Afterwards they were heated to 200 °C at a rate of 5 °C/min and held isothermally for 3 min

before cooling to -70 °C at a rate of 5 °C/min. The samples were finally reheated to 200 °C at a

rate of 5 °C/min. The % crystallinity of the polymers, Xc, was estimated using equation (4.2).

100H

HX

100

mc

(4.2)

where ΔHm is the enthalpy of fusion and ΔH100 is the theoretical fusion enthalpy of a 100%

crystalline polymer, which is 146 J/g [74].

Page 57: Preparation and characterization of biopolymer compounds

39

4.2.4.2 Thermogravimetric analysis (TGA)

TGA was conducted on a Q500 TGA from TA Instruments under nitrogen atmosphere, using 6-8

mg samples. The weight loss was evaluated by heating to 800 °C at a heating rate of 20 °C/min.

Isothermal experiments were conducted at the compounding temperature of 190 °C for 20 min.

The temperature at which 5 wt.% degradation occurred was reported as the degradation onset

temperature.

4.2.4.3 Scanning electron microscopy

Blend morphologies were observed using a JEOL JSM-840 scanning electron microscope.

Samples were first hot-pressed at 200oC for 3 min, then immersed in liquid nitrogen for 3 min

before brittle fracture. The MCL PHA phase was etched in acetone overnight at room

temperature.

4.2.4.4 Rheology

Compression molded discs, 25 mm diameter and 2 mm thick, were prepared using a Carver

press. The linear viscoelastic properties were measured in the oscillatory mode using a stress

controlled rheometer, Visco Tech from Reologica, under nitrogen purge. Frequency sweeps

were conducted at 190 °C using a cone and plate fixture having 25 mm diameter and 2° angle,

at a frequency range between 1-100 rad/s. This range was chosen to limit the duration of the

experiment to 100 s, given the sensitivity of PHB to degradation. Time sweeps were conducted

at a frequency of 6.28 rad/s and temperatures ranging from 135 to 155 °C to obtain cure curves

for the cross-linked PHO formulations.

Page 58: Preparation and characterization of biopolymer compounds

40

4.2.4.5 Mechanical properties

The compounded materials were pre-dried in a vacuum oven at room temperature overnight.

Specimens for mechanical property characterization were prepared by compression molding

using a Carver press under 5000 N force, at 200 °C and a residence time of 3 min, then

quenched in cold water. All specimens were conditioned at room temperature for 48 h after

compression molding, prior to mechanical testing. Tensile tests were conducted in accordance

with ASTM D638 using standard type V test specimens, with an Instron 3369 Universal tester, at

a cross head speed (CHS) of 5 mm/min. The average of five measurements is reported. Un-

notched Izod impact tests were conducted in accordance with ISO 180 using standard

specimens on a SATEC Instron machine and the average of five specimens are also reported.

4.3 Results and discussion

4.3.1 Thermal and rheological properties

Thermal degradation is an important concern when processing PHB. The isothermal TGA curves

(Figure 4.1) show that under nitrogen atmosphere PHB started to degrade at times longer than

10 min at the compounding temperature of 190 °C. Based on the TGA data, PHO had better

thermal stability than PHB and its addition to PHB improved the thermal stability of the blend.

The by-product of PHB degradation is mainly trans-2-butenoic acid [86], whereas the products

of PHO degradation would be a mixture of trans-2-octenoic and trans-2-hexenoic acids. The

presence of different decomposition products may affect the degradation kinetics of the

mixture.

Page 59: Preparation and characterization of biopolymer compounds

41

Figure 4.1 TGA curves for PHO, PHB and the 85/15 PHB/PHO blend at 190 °C

Table 4.1 summarizes the degradation onset temperatures of the blends in non-isothermal

experiments. Addition of PHO gradually increased the degradation onset temperature of the

blends. The shift in temperature was as high as 20 °C for the blends containing 30 wt.% PHO.

Figure 4.2 shows the DSC thermograms of PHB and PHB/PHO blends. The first heating cycle of

the PHB showed a single melting peak at 181 °C and a corresponding crystallinity of 66%. A

double peak appeared during the second heating cycle (Figure 4.2a). The first melting peak at

the lower temperature (171 °C) corresponds to crystals formed at the crystallization

temperature and the second one at 177 °C is due to crystals that form during the heating cycle

[83,87,88]. The upper peak is generally attributed to the presence of more stable crystals that

94

95

96

97

98

99

100

101

0 5 10 15 20

We

igh

t (%

)

Time (min)

PHO

PHB

85/15 PHB/PHO

Page 60: Preparation and characterization of biopolymer compounds

42

are favoured when unstable crystals melt and reorganize during the heating scan at slow

cooling rates, such as the ones employed in DSC [88].

Table 4.1 Crystallization temperature (TC), first and second melting peaks (TM1 and TM2

respectively), % crystallinity and degradation onset temperature for PHB and PHB/PHO blends.

PHB/PHO

(wt.%/ wt.%) TC (°C) TM1 (°C) TM2 (°C) Crystallinity (%)

Degradation

onset temp (°C)

100/0 101 171 177 64 254

95/5 97 171 177 58 261

90/10 91 173 178 55 261

85/15 96 172 178 53 264

80/20 87 172 177 53 273

70/30 78 169 177 45 277

The relative magnitude and position of these peaks generally depends on the heating rate [89],

and the blend composition [23]. In the present study, the high temperature endotherm became

more prominent as the PHO content increased, suggesting that crystals that formed during the

heating cycle were favored at higher PHO content. Moreover, as the amount of PHO was

increased in the blend, the PHB crystallization peak shifted to lower temperatures (Figure 4.2b),

suggesting that the PHB crystalline structure is affected in the presence of PHO.

Page 61: Preparation and characterization of biopolymer compounds

43

Figure 4.2 a) DSC endotherm (2nd heating cycle) and b) DSC exotherm of PHB and PHB/PHO

blends

150 155 160 165 170 175 180 185

Heat

Flo

w (

W/g

)

Temperature ( C)

30 50 70 90 110 130 150

He

at

Flo

w (

W/g

)

Temperature ( C)

5 wt % PHO 15 wt % PHO

30 wt % PHO PHB

a)

b)

Page 62: Preparation and characterization of biopolymer compounds

44

The PHO had very low crystallinity. A first heating cycle revealed a Tg of -36 °C, melting

temperature of 63 °C and crystallinity of 15 %. Given its very low crystallization rates, PHO did

not crystallize during the cooling cycle, therefore the second heating cycle was entirely

featureless and is not shown in Figure 4.2. Addition of PHO to PHB resulted in a gradual

decrease in the heat of fusion, resulting in decreased crystallinity, as shown in Table 4.1. This is

a common phenomenon when an elastomeric material is added to a semi-crystalline one, and

has also been noted when PHB was blended with amorphous polymers [81].

The linear viscoelastic properties of PHB and PHO are summarized in Figure 4.3. Both polymers

exhibited Newtonian behavior, with no shear thinning over the frequency range investigated, as

shown in Figure 4.3a. A large viscosity mismatch between the two polymers at the

compounding temperature was noted, with PHB being significantly more viscous. This is

obviously attributed to the differences in the molecular weight (172,000 Da and 1,416,000 Da

for PHO and PHB respectively). The viscosity ratio, ⁄ , (where ηd is the viscosity of the

dispersed phase and ηm is the viscosity of the matrix) is very low, about 0.03 and affects

negatively the dispersion of the dispersed phase within the matrix, as discussed in the following

section.

4.3.2 Morphology

The PHB/PHO blends had droplet-matrix morphology, typical of immiscible polymer blends, as

seen in the SEM images (Figure 4.4). The PHO domains were well dispersed within the PHB

matrix at low PHO content, but as the level of PHO was increased, the domain sizes became

larger and the morphology deteriorated. At 30 wt.% PHO, the coalescence of the dispersed

Page 63: Preparation and characterization of biopolymer compounds

45

Figure 4.3 Rheological properties of PHB and PHO and effect of peroxide cross-linking on a)

complex viscosity, b) storage modulus and c) loss tangent, tan δ, measured at 190 °C

1

10

100

1000

10000

100000

1 10 100

Co

mp

lex

Vis

co

sit

y (

Pa

s)

Frequency (rad/s)

100

102

103

100 101 102

104

105

101

0.01

0.1

1

10

100

1000

10000

100000

1 10 100

Sto

rag

e M

od

ulu

s

(Pa

)

Frequency (rad/s)

102

103

104

101

100

10-1

10-2

100 101 102

105

0.01

0.1

1

10

100

1000

1 10 100

tan

δ

Frequency (rad/s)

PHB PHOPHO-0.06 wt % L-231 PHO-0.2 wt % L-231PHO-0.5 wt % L-231

10-2

10-1

100

101

102

103

100 101 102

a)

b)

c)

Page 64: Preparation and characterization of biopolymer compounds

46

phase resulted in a very coarse structure. As explained above, the tendency for coalescence

may be attributed to the significant viscosity mismatch between the two blend components.

Better morphology was reported in solution blending of PLA and PHO having similar viscosities

[9]. However viscosity is not a factor during solution blending, therefore the results are not

directly comparable.

Figure 4.4 Scanning electron microscopy blends containing a) 5 wt.%, b) 10 wt.%, c) 15 wt.%, d)

20 wt.% and e) 30 wt.% of PHO at 2000x magnification

4.3.3 Mechanical properties

The tensile strain increased with the PHO content (Figure 4.5a). Considerable enhancement was

observed above 15 wt.% PHO, indicating improved flexibility of the blends. At 30 wt.%, there

was a decline in tensile strain. At this high PHO content, the PHO domains coalesced (Figure

4.4) and the deterioration in morphology caused the tensile strain to decrease. The un-notched

impact strength of the blends improved by only 50% when 20 wt.% PHO was added, but

Page 65: Preparation and characterization of biopolymer compounds

47

increased by 150% with 30 wt.% PHO compared to pristine PHB. On the other hand, the tensile

stress and the Young’s moduli decreased, as shown in Figure 4.5b. The decrease in tensile stress

is typical of impact modification and can be justified by the decrease in crystallinity with the

addition of PHO in the blend and the introduction of a softer component in the blend.

Improvements in the strain at break, and a decrease in Young’s modulus and tensile strength

have been reported in PHB-HV/MCL PHA blends prepared by solution blending [84]. However in

these blends MCL PHA contents above 5 wt.% led to phase separation and a decrease in strain

at break. The results outlined above suggest that addition of PHO to PHB reduced the

crystallinity of the blend, and moderately increased impact and elongation, which were

counteracted by a decrease in the modulus. These are the expected results of impact

modification. The extent of impact modification however remains limited, due to the coarse

morphology, especially at high loadings which is attributed to the viscosity mismatch between

the blend components, attributed to the very low viscosity of PHO.

4.3.4 Cross-linking of PHO

Chain extension of PHO through chemical cross-linking was employed, in an attempt to increase

the viscosity of the dispersed phase to attain a more favourable viscosity ratio. Cross-linking of

MCL PHAs can be achieved using peroxides, radiation, or sulfur cures. Gagnon et al. [90] cross-

linked saturated and unsaturated MCL PHAs using four different types of peroxide with and

without coagents. They found that cross-linking decreased the crystallinity of the polymer.

Reduced tensile and tear strength was observed as a result of chain scission. Sulfur

vulcanization was also used by Gagnon et al. [17], whereas Dufresne et al. [91] cross-linked MCL

PHA by irradiation.

Page 66: Preparation and characterization of biopolymer compounds

48

Figure 4.5 a) Un-notched impact strength and tensile strain b) tensile stress and Young’s

modulus of pristine PHB and its blends

0

2

4

6

8

10

12

14

0

2

4

6

8

10

12

14

16

0% 5% 10% 15% 20% 30%

Ten

sil

e S

train

(%

)

Un

no

tch

ed

Im

pact

(KJ/m

2)

PHO in PHB

Unnotched Impact

Tensile Strain

0

100

200

300

400

500

600

700

800

0

5

10

15

20

25

30

0% 5% 10% 15% 20% 30%

Yo

un

g's

Mo

du

lus

(M

Pa

)

Te

ns

ile

Str

es

s (

MP

a)

PHO in PHB

Tensile stress (MPa)

Young's Modulus(MPa)

a)

b)

Page 67: Preparation and characterization of biopolymer compounds

49

In this work, given the sensitivity of the polymers to temperature, lauroyl peroxide was used,

because it decomposes at relatively low temperatures. A series of time sweeps were conducted

at various temperatures using the rheometer to generate a series of cure curves shown in

Figure 4.6. Based on these data, 155 °C was chosen for the cross-linking reaction, aiming for the

shortest possible reaction time.

Figures 4.3a-c show the effect of cross-linking on the linear viscoelastic properties of PHO.

Addition of peroxide increased significantly the complex viscosity and storage modulus of PHO,

whereas the loss tangent, tanδ, decreased below 1, revealing a transformation from viscoelastic

liquid to a viscoelastic solid.

Figure 4.6 Cure curves of PHO with 0.1 wt.% lauroyl peroxide as a function of temperature at a

frequency of 1 Hz.

10

100

1000

10000

0 500 1000 1500 2000

Sto

rag

e M

od

ulu

s (

Pa

)

Time (s)

135 ° C

145 ° C

155 ° C

Page 68: Preparation and characterization of biopolymer compounds

50

A peroxide content of 0.06 wt.% was sufficient to achieve 67 wt.% gel content in the cross-

linked PHO, resulting in a tanδ value of about 1. Further increases in peroxide to 0.2 and 0.5

wt.% resulted in almost fully gelled material with gel contents of 87 and 97% respectively.

It should be noted that cross-linking resulted in a significant drop in the crystallinity of PHO, as

recorded from the 1st heating cycle, from 15 to 7% with 0.2 wt.% Lauroyl peroxide, whereas the

fully cross-linked PHO was completely amorphous.

The increases in viscosity upon cross-linking were accompanied by a substantial increase in

shear thinning behavior, and loss of the Newtonian plateau, as expected for cross-linked

polymers having high cross-link densities. Given the change in the viscosity-shear rate

dependence, cross-linked PHA can only match the viscosity of PHB in a very narrow

frequency/shear rate range. Based on the data of Figure 4.3, in order to match the viscosity of

PHO in the shear rate range of 10-100 s-1, which is relevant to compounding, 0.2-0.5 wt.% of

peroxide is needed.

The morphology of the blends containing 30 wt.% PHO cross-linked with different amounts of

lauroyl peroxide is compared to that of the unreacted blends in Figure 4.7. The corresponding

viscosity ratios, calculated at a representative shear rate of 50 s-1 from Figure 4.3a are shown in

the caption of Figure 4.7. The dispersed PHO domain size became progressively smaller upon

increasing the amount of cross-linking (Figure 4.7b and 4.7c). Significant improvement in

morphology was seen upon addition of 0.2 wt.% Lauroyl peroxide. It was impossible to assess

the domain size in the blends containing PHO cross-linked with 0.5 wt.% Lauroyl peroxide,

because the high gel content did not allow for etching and thus sufficient contrast (Figure 4.7d).

Page 69: Preparation and characterization of biopolymer compounds

51

The improvement in morphology correlates well with the decrease in the viscosity ratio

achieved by using the cross-linked PHO dispersed phase.

Figure 4.7 SEM of blends of PHB/cross-linked PHO 70/30 blends a) uncross-linked (viscosity

ratio, λ=0.03) b) cross-linked with 0.06 wt.% of lauroyl peroxide (λ=0.15) c) cross-linked with 0.2

wt.% of lauroyl peroxide (λ=0.36) d) cross-linked with 0.5 wt.% of lauroyl peroxide (λ=3.73).

As shown in Table 4.2, there were significant improvements in the Young’s modulus and strain

at break of the blends containing cross-linked PHO, but the impact strength was unchanged.

This may be due to the high cross-link densities and gel content of the cross-linked PHO, which

alter its thermoplastic elastomer nature.

a) b)

)

a)

c)

)

b

a)

d)

)

b

a)

Page 70: Preparation and characterization of biopolymer compounds

52

Table 4.2 Comparison of mechanical properties of PHB/PHO 70/30 blends containing uncross-

linked and cross-linked PHO

These results suggest that matching the viscosity of the blend components can result in finer

morphology and improvements in the mechanical properties. However high cross-link densities

and gel content of the cross-linked PHO, may present a limitation in terms of impact properties.

Therefore tight control of the morphology, while avoiding the formation of excessive gels, is

crucial to achieve desirable property improvements.

4.4 Conclusions

PHB/PHO blends had improved thermal stability, tensile strain at break and unnotched impact

strength compared to the unmodified PHB. The ability of PHO to act as an impact modifier for

PHB was limited by the viscosity mismatch between the two components, which resulted in a

coarse blend morphology. Chain extension of PHO by peroxide cross-linking improved the

PHB/PHO Tensile stress

(MPa)

Tensile strain

at break (%)

Young's modulus

(MPa)

Un-notched

impact (KJ/m2)

70/30 11.0 (±2.0) 9.6 (±0.5) 189 (±23) 14.2 (±0.4)

70/30 (0.2 wt.%

lauroyl peroxide ) 10 (±1.6) 14.4 (±0.3) 220 (±20) 13.3 (±1.4)

70/30 (0.5 wt.%

lauroyl peroxide ) 10.6 (±0.8) 14.5 (±0.4) 243 (±27) 14.2 (±0.5)

Page 71: Preparation and characterization of biopolymer compounds

53

viscosity of PHO and led to better morphology and improved modulus and elongation at break

of the blends.

Page 72: Preparation and characterization of biopolymer compounds

54

Chapter 5 Dramatic improvements in strain hardening and

crystallization kinetics of PLA by simple reactive modification in the

melt state*

5.1 Introduction

Poly(lactic acid)(PLA) is a bioderived, biodegradable thermoplastic polyester [92], which can be

processed using conventional thermoplastics processing equipment, including injection

molding, blow molding, film casting and blowing [5]. However it has a very narrow processing

window, because of the lack of melt strength and its slow crystallization rates. Additionally its

poor engineering properties, including impact strength and heat resistance have mainly

confined its applications to food packaging [38], as well as biomedical applications, such as drug

delivery, where biocompatibility and biodegradability are desired [93]. The properties of PLA

depend significantly upon its molecular weight and the stereochemical makeup of the

backbone, which is controlled by polymerization with D-lactide, L-lactide, or D,L-lactide, to form

random or block stereocopolymers [37]. Minimizing the amount of D-lactide is required to

obtain PLA with higher crystallinity; however most commercial grades have low crystallinities

and low crystallization rates, unless nucleating agents are used. The rheological properties of

PLA depend on the molecular weight and molecular weight distributions, presence of

branching, as well as its stereochemical makeup [94-98]. Strain hardening has been reported in

melts containing a high molecular weight tail [94], and in amorphous PLA containing mixtures

of the D and L isomers [96] at low temperatures, but otherwise it is generally accepted that

commercially available linear PLA lacks the level of strain hardening, and therefore melt

*A version of this chapter is accepted. Nerkar, M., Ramsay, J.A., Ramsay, B.A., Kontopoulou, M. Macromolecular Materials and Engineering, accepted May2014.

Page 73: Preparation and characterization of biopolymer compounds

55

strength, needed for normal processing operations. This restricts its processability in operations

involving high stretch rates, such as film blowing, thermoforming, foaming etc. Given these

shortcomings, approaches have been proposed to achieve chain extension and/or branching in

PLA, with branching generally considered more beneficial [99]. Compared to the various

synthetic routes that exist to synthesize branched PLA, methods that employ reactive

modifications in the melt state are generally considered to be more convenient and industrially

relevant. Various modification approaches to improve processability have been summarized by

Pilla et al. [100] and Yu et al. [101]. These include chain extension in the presence of glycidol

[46] and long chain branching via functional group reactions of pyromellitic dianhydride and

triglycidyl isocyanurate [99]. Furthermore chain extenders, such as tris (nonylphenyl)

phosphate, polycarbodiimide and multi-functional epoxy compounds have been used to

counteract degradation in polyesters, such as PLA and to achieve chain extension

[47,51,102,103].

Reactive extrusion of PLA using organic peroxides has been undertaken to increase the

molecular weight, viscosity and melt strength with limited success, as the resulting branching is

often counter-balanced by severe chain scission [52-54]. Radiation induced cross-linking in the

presence of multi-functional coagents has been suggested as an alternative, but generally

resulted in physical property reduction [104].

Peroxide-initiated reactive extrusion in the melt state, assisted by coagents is frequently

employed as a means to introduce long-chain branching in linear polymers, such as

polypropylene[105,106]. However there are only two reports, employing this approach in PLA.

Yang et al. [107] used triallyl isocyanurate as a cross-linking agent together with dicumyl

Page 74: Preparation and characterization of biopolymer compounds

56

peroxide (DCP), to obtain compounds with different levels of cross-linking. More recently, You

et al. [108] reported that PLA prepared through reaction with DCP and pentaerythritol

triacrylate (PETA) coagent had enhanced viscoelastic properties, which was attributed to

branching. The resulting product had faster crystallization rates under isothermal conditions.

However in these publications there was no mention about the properties of the resulting

materials under uniaxial extension and the non-isothermal crystallization behavior of the

polymers, which is relevant to processing, was not reported.

This chapter reports substantial improvements in the melt strength and non-isothermal

crystallization kinetics, upon employing a simple chemical modification method in the melt state

using solvent-free, peroxide-initiated grafting of a multi-functional coagent (triallyl trimesate,

TAM). To the best of our knowledge this is the first time that simultaneous improvements in all

these properties upon reactive modification are reported for PLA. These attributes are expected

to enable use of these materials in operations such as foaming, injection moulding and film

processing.

5.2 Experimental

PLA (grade 3251D, MFI 35 g/10 min at 190 °C/ 2.16 kg) was obtained from Natureworks®. TAM

(98%, Monomer Polymer Inc.), dicumyl peroxide (DCP, 98%, Sigma-Aldrich), acetone (Sigma-

Aldrich) and tetrahydrofuran (THF, Sigma-Aldrich) were used as received. Joncryl® ADR 4368 a

multi-functional epoxide styrene-acrylic oligomeric chain extender, containing glycidyl

methacrylate (GMA) functions was supplied by BASF. It has a functionality of 9 with epoxy

equivalent weight 285 g/mol, and molecular weight 6,800 g/mol [51].

Page 75: Preparation and characterization of biopolymer compounds

57

PLA was dried in a vacuum oven at 100 °C for 3 h to remove moisture. Peroxide-degraded PLA

(PLA/DCP) was prepared by coating ground PLA powder (15 g) with an acetone solution

containing DCP (0.045 g) and allowing the solvent to evaporate. The resulting mixture,

containing 0.3 wt.% DCP, was charged to a DSM micro-compounder, equipped with twin-co-

rotating screws, at 180 °C at 100 rpm for 6 min. Coagent-modified PLA was prepared as

described for PLA/DCP from a mixture of PLA (14.8 g), DCP (0.045 g) and TAM (0.15 g), yielding

PLA/TAM containing 0.3 wt.% DCP and 1 wt.% TAM. A compound containing 1.2 wt.% of GMA,

which was the amount required to yield similar zero shear viscosity as the PLA/TAM was also

prepared and used for comparison. Neat PLA was processed under the same conditions

outlined above, to provide a suitable basis for comparison. After compounding, the strands

were quenched in cold water before chopping into pellets for further characterization.

Samples were prepared for size exclusion chromatography (SEC) analysis by dissolving 10 mg of

polymer in 1 mL of distilled THF overnight to ensure complete dissolution, and filtered through

a 0.2 µm nylon filter. Polymer molecular weight distributions (MWD) were determined with

respect to polystyrene standards using a Viscotek 270max separation module with triple

detection by differential refractive index (DRI), viscosity (IV) and light scattering (low angle LALS

and right angle RALS), which was maintained at 40 °C and contained two porous PolyAnalytik

columns in series. Distilled THF was used as the eluent at a flow rate of 1 mL/min.

The linear viscoelastic properties were measured in the oscillatory mode using a stress

controlled rheometer (Visco Tech by Reologica). Frequency sweeps were conducted at 180 °C

using 20 mm parallel plates, under nitrogen purge. Samples were further characterized in

uniaxial extension using an SER-2 universal testing platform from Xpansion Instruments hosted

Page 76: Preparation and characterization of biopolymer compounds

58

on an MCR-301 Anton Paar rheometer. Measurements were conducted at 180 °C at Hencky

strain rates ranging from 0.10 to 10 s-1. The linear viscoelastic (LVE) oscillatory measurements

obtained at 180 °C were used to calculate the LVE stress growth curve, η+, and to check the

consistency of the extensional measurements. The curve corresponding to 3η+ represents the

LVE envelope in uniaxial extension, according to Trouton’s law.

Differential scanning calorimetry (DSC) was conducted using a DSC Q 100 by TA Instruments.

Samples were scanned between 0 and 200 °C at a heating rate of 5 °C/min. After the first

heating, each sample was held isothermally at 200 °C for 3 min before cooling at rates between

2.5 and 20 °C/min, to determine the crystallization onset and peak temperatures according to

ASTM D3418. The % crystallinity of PLA, was estimated using Equation (5.1)

100100

H

HHX ccm

c (5.1)

where ΔHm is the apparent fusion enthalpy, ΔHcc is the exothermic enthalpy that is absorbed by

crystals formed during the heating scan and ΔH100 is the theoretical fusion enthalpy of a 100%

crystalline polymer, which is 93.6 J/g for PLA [109].

Isothermal studies involved heating the sample to 200 °C and holding it for 3 min, followed by

cooling at 50 °C/min to temperatures ranging from 135-155 °C, where they were held

isothermally until completion of crystallization. The analysis included evaluations of the relative

crystallinity as a function of time and standard Avrami kinetics [110].

Page 77: Preparation and characterization of biopolymer compounds

59

5.3 Results and Discussion

5.3.1 Rheological characterization

The linear viscoelastic properties of neat PLA, PLA reacted with 0.3 wt.% DCP (PLA/DCP) and

with 0.3 wt.% DCP/ 1 wt.% TAM (PLA/TAM) are summarized in Figure 5.1. PLA/DCP had

essentially unaltered flow characteristics compared to the unmodified PLA (Figure 5.1a) and

linear architecture, as revealed by the Van Gurp-Palmen plots of Figure 5.1b, tend to the limit of

90°. This suggests that peroxide-induced degradation was compensated by chain extension,

without branching. This is corroborated by the minimal differences in the molar mass

distributions between these two compounds (Table 5.1).

On the contrary, PLA/TAM demonstrated a substantial increase in molar mass (Table 5. 1) and

melt viscosity (Figure 5.1a). It is well-known that solvent-free processing with multi-functional

coagents involves simultaneous chain scission and cross-linking, the balance of which controls

the molecular weight and branching distributions of the final product. In general, reaction with

coagents bearing multiple acrylate, allylic or styrenic groups results in bimodal molecular

weight and branched distribution, comprised of a linear chain population of relatively low

molecular weight, and a high molecular weight hyper-branched chain population, which can

progress above the gel point [56,111]. Furthermore it has been reported that systems

containing polypropylene reacted with coagent and peroxide can undergo a precipitation

polymerization, which results in the formation of a separate phase of highly cross-linked,

coagent-rich sub-micron sized particles [56,112]. The resulting products possess a creamy

appearance in the melt state, owing to the presence of these cross-linked nano-particles [56].

Page 78: Preparation and characterization of biopolymer compounds

60

Figure 5.1 a) Complex viscosity as a function of frequency and b) phase degree as a function of

complex modulus at 180 °C.

In the present case, treatment with 0.3 wt.% DCP and 0.1 wt.% TAM produced a creamy (as

opposed to the transparent PLA and PLA/DCP), gel-free product with increased melt elasticity,

and shear thinning, as observed in Figure 5.1, consistent with the presence of branching [57].

We compared the properties of PLA/TAM to those of PLA chain extended using a multi-

100

1000

10000

0.1 1 10 100 1000

Co

mp

lex

Vis

co

sit

y (

Pa

s)

Frequency (rad/s)

40

50

60

70

80

90

100

100 1000 10000 100000

Ph

as

e A

ng

le ( )

Complex Modulus (Pa)

PLA

PLA/DCP

PLA/TAM

PLA/GMA

b)

)

b

a)

a)

)

b

a)

Page 79: Preparation and characterization of biopolymer compounds

61

functional epoxide styrene-acrylic oligomeric chain extender, containing GMA functions (trade

name Joncryl® from BASF).

Table 5.1 Material characterization

Material Mwa)

[g/mol]

PDIb) TM

[°C]

TC

[°C]

TC,onset

[°C]

TCC

[°C]

Crystallinity

[%]

PLA 98,140 1.7 173 N/A N/A 109 24

PLA/DCP 90,600 1.8 170 N/A N/A 94 35

PLA/TAM 143,020 2.0 169 133 149c), 142d) N/A 52

PLA/GMA 115,000 1.7 168 105 124c), 123d) 95 34

a)Mw: Weight average molar mass; b)PDI: polydispersity index; c)at cooling rate of 2.5 °C, d)at cooling rate of 5 °C.

The chain extended (PLA/GMA) was prepared by reacting PLA with 1.2 wt.% Joncryl®, which

was the amount needed to match the zero shear viscosity of PLA/TAM. The epoxy functions

contained within multi-functional epoxies can react with the –OH and –COOH end groups of

PLA, resulting in random branching and/or cross-linking [51]. The low levels of Joncryl® used

herein, produced a gel-free product with higher molar mass than the parent PLA (Table 5.1)

increased viscosity and deviations from the terminal flow behavior (Figure 5.1).

The different shear thinning characteristics and shapes of the Van Gurp-Palmen plots of

PLA/GMA compared to PLA/TAM point to different branching levels and chain topologies.

PLA/TAM presumably contained small amounts of a hyper-branched population, which are

characteristic of this coagent modification, whereas reaction with multi-functional epoxies

Page 80: Preparation and characterization of biopolymer compounds

62

produces random branching. In spite of the different mechanisms of chain extension/branching,

the ultimate strain hardening characteristics of PLA/TAM and PLA/GMA were very similar, as

shown in Figure 5.2. Pronounced strain hardening was present in both materials, providing

evidence of long chain branching. Note that the viscosities of the parent PLA and PLA/DCP were

below the threshold needed to support extensional viscosity experiments.

Figure 5.2 Tensile stress growth coefficient (ηE+) of TAM and GMA modified PLA as a function of

strain rate and time at Hencky strain rates of 0.1, 1 and 10 s-1 at 180 °C. Curves are shifted by an

arbitrary factor for the sake of clarity. Solid lines represent the LVE envelop (3E+) for each

sample.

Page 81: Preparation and characterization of biopolymer compounds

63

5.3.2 Thermal properties

Detailed DSC data are presented in Table 5.1 and Figure 5.3. The glass transition temperature

(Tg) of PLA was 62 °C and remained unchanged in all modified materials. PLA had a cold

crystallization peak, Tcc, at 109 °C, a melting peak, TM, at 173 °C (Table 5.1 and Figure 5.3a).

PLA/DCP had reduced cold crystallization temperature and increased crystallinity. These findings

are commonly associated to PLA degradation [51,108]. Neither of these two materials

crystallized during the cooling cycle.

Figure 5.3 DSC a) 2nd heating scan at rate of 5 °C/min b) cooling scan at the rate of 5 °C/min

70 90 110 130 150 170 190

Hea

t F

low

A

.U.

Temperature ( C)

PLA

PLA/GMA

PLA/TAM

70 90 110 130 150

He

at

Flo

w A

.U.

Temperature ( C)

PLA/DCP

PLA

PLA/GMA

PLA/TAM

PLA/DCP

a)

b)

Page 82: Preparation and characterization of biopolymer compounds

64

On the contrary, the branched PLAs showed exothermal crystallization peaks. PLA/GMA had a

weak crystallization peak, TC, around 105 °C, suggesting a moderate effect of this modification

on the ability of the chains to crystallize. On the other hand, coagent modified PLA had a clear

and sharp crystallization peak at 133 °C (Figure 5.3b). This was accompanied by the

disappearance of the cold crystallization peak, and a significant increase in crystallinity by 117 %

with respect to neat PLA and 50 % with respect to PLA/DCP. Even though changes in the cold

crystallization of PLA upon modification with a PETA coagent have been mentioned previously

[108] this is the first time that the presence of an exothermic crystallization peak arising during

the cooling cycle, which is indicative of the capability of the material to crystallize upon cooling

during normal polymer processing operations, is reported.

The ability of our modified materials to crystallize was evident not only by the appearance of an

exothermic crystallization peak, but also by their isothermal and non-isothermal crystallization

kinetics (Figure 5.4). The PLA and PLA/DCP formulations did not crystallize and therefore are

not included in this comparison. Plots of the evolution of relative crystallinity as a function of

time revealed a crystallization half-time (t1/2) of 9.3 min at 135 °C for PLA/GMA, whereas the

half-time of PLA/TAM at this temperature was only 0.6 min. The results of the Avrami analysis

for temperatures ranging from 135-155 °C are presented in Table 5.2. The Avrami exponents

suggest similar crystal growth habit in all cases. Introduction of branching in polymers such as

polypropylene has been associated previously to changes in the crystallization kinetics [113].

The relative crystallinity as a function of time, recorded during non-isothermal crystallization

experiments is shown in Figure 5.4b. While cooling from the melt state, PLA/TAM started to

Page 83: Preparation and characterization of biopolymer compounds

65

crystallize significantly earlier, with t1/2 of 2.6 and 1.8 min at cooling rates of 2.5 and 5°C/min

respectively, as compared to 8.4 and 3.9 min for PLA/GMA.

Figure 5.4 Relative degree of crystallinity as a function of time a) isothermal crystallization

experiments; (-) PLA/TAM at 135 °C, ()PLA/TAM at 140 °C, ()PLA/TAM at 150 °C, (◆)

PLA/GMA at 135 °C and (b) non-isothermal crystallization experiments; ()PLA/TAM at 2.5

°C/min, (◆)PLA/TAM at 5 °C/min, (o)PLA/TAM at 20 °C/min, ()PLA/GMA at 2.5 °C/min,

()PLA/GMA at 5 °C/min

0

25

50

75

100

0 2 4 6 8 10

Rela

tive

de

gre

e o

f c

rys

tall

init

y (

%)

Time (Min)

0

25

50

75

100

0 5 10 15 20

Re

lati

ve

de

gre

e o

f c

rys

tall

init

y (

%)

Time (Min)

a)

b)

Page 84: Preparation and characterization of biopolymer compounds

66

Furthermore PLA/TAM had t1/2 of 0.96, 0.73 and 0.66 min at cooling rates of 10, 15 and 20 °C

respectively, whereas PLA/GMA did not crystallize at these conditions.

Table 5.2 Isothermal Avrami constants and crystallization half time for PLA/GMA and PLA/TAM

at various temperatures

Temperature [°C] n K [min-1] t1/2 [min]

PLA/GMA 135 2.9 0.0001 9.3

PLA/TAM 135 3.0 4.27 0.6

PLA/TAM 145 3.4 0.14 1.6

PLA/TAM 150 3.3 0.02 2.8

PLA/TAM 155 2.6 0.01 6.1

The findings reported above point to a nucleating effect, which occurred in spite of the absence

of a nucleating agent. As explained earlier, a nucleating effect attributed to the formation of a

separate phase of coagent-rich particles that forms upon reactive modification as a result of

TAM oligomerization [55], was reported recently in coagent modified polypropylene. We

suggest that a similar nucleation effect is responsible for the enhancements in crystallinity and

crystallization rates in the reactively modified PLA/TAM product.

5.4 Conclusions

PLA with long chain branching was produced by a simple radical mediated peroxide-initiated

grafting of TAM coagent in the melt state. The resulting product demonstrated strain hardening,

Page 85: Preparation and characterization of biopolymer compounds

67

consistent with its long chain branched characteristics, and significantly enhanced crystallinity

and crystallization rates, suggesting that this is a promising approach to enhance the processing

characteristics and properties of PLA.

Page 86: Preparation and characterization of biopolymer compounds

68

Chapter 6 Improvements in the extensional rheology, thermal

properties and morphology of poly(lactic acid)/ poly-3-

hydroxyoctanoate blends through reactive modification

6.1 Introduction

Among the key challenges associated with more wide-spread acceptance of biopolyesters, such

as poly(lactic acid) (PLA) and polyhydroxyalkanoates (PHA) in engineering applications are their

high production costs, brittle nature, slow crystallization rates and low melt strength, which

restrict their processability under common polymer processing operations, as well as their

hygroscopic nature, and susceptibility to degradation.

The advantages and drawbacks of PLA and PHAs as well as the current state-of-the-art of the

various modification methods that have been employed to overcome their limitations have

been reviewed recently [24,36,38,79,104,114]. Blending with ductile polymers and addition of

plasticizers are commonly used to improve the properties and processability of PLA [36,114],

and poly-3-hydroxybutyrate (PHB) [79,115]. Furthermore, to achieve fully bio-based

formulations, blends of biopolymers have been studied extensively [79,80,116].

PHAs and their copolymers have been used to enhance the toughness of PLA through solution

blending [117] or melt blending [117-119]. As already shown in Chapter 4, medium-chain-

length (MCL) PHAs can serve as impact modifiers for PHB, due to their low crystallinity and

elastomeric character [84,100,101].

To address the problem of lack of melt strength and suitable rheological properties of

biopolyesters, reactive modifications in the melt state have been employed to achieve chain

extension. Various modification approaches of PLA have been summarized by Pilla et al. [100],

Page 87: Preparation and characterization of biopolymer compounds

69

and Yu et al. [101]. These include chain extension of PLA in the presence of glycidol [46] and

introduction of long chain branching via functional group reactions of pyromellitic dianhydride

and triglycidyl isocyanurate [99]. Furthermore chain extenders, such as tris (nonylphenyl)

phosphite, polycarbodiimide and epoxy-functionalized oligomeric acrylic copolymer (trade

name Joncryl® from BASF) have been considered to counteract degradation in PLA and

introduce chain extension [47,102,103]. In-situ cross-linked hyperbranched polymers have been

used to improve the toughness of PLA [120,121]. Reactive extrusion of PLA using organic

peroxides and coagents has also been undertaken [52-54,104,107,108]. On the other hand, as

discussed in Chapter 4, cross-linking of MCL PHAs can be achieved using peroxides, radiation, or

sulfur cures [17,90,91].

Reactive blending of biopolyesters, such as PHB and polyhydroxybutyrate-co-valerate (PHBV)

with polybutylene succinate (PBS) [31], and PLA with polycaprolactone (PCL) [122], PLA with

PBS [123], PLA with Poly(butylene adipate-co-terephthalate) (PBAT) [124] and PHB [125] using

peroxides has been employed to produce in-situ compatibilized blends having improved

properties. Another proposed chemical modification involved blending of PLA with MCL-PHAs

using diisocyanate chain extenders, which are highly toxic compounds [4]. Epoxy functionalities

have been introduced in MCL PHA to react with the hydroxyl groups of PLA, thus increasing the

interfacial interaction and improving the blend morphology and compatibility [117]. This

approach was also tested in the present thesis (Appendix A) with limited success.

In chapter 5, we presented a simple reactive modification approach, utilizing solvent-free,

peroxide-initiated grafting of a multi-functional coagent, to introduce branching and achieve

substantial improvements in the strain hardening characteristics of PLA. This approach resulted

Page 88: Preparation and characterization of biopolymer compounds

70

in faster crystallization kinetics, both under isothermal and under non-isothermal conditions.

The present chapter investigates chain extension of PHO and PLA/PHO blends using the same

approach. The properties of the reactively modified blends are compared to those of

unmodified blends.

6.2 Experimental

6.2.1 Materials

Polyhydroxyoctanoate (PHO) containing ~98 mol% of 3-hydroxyoctanoate and ~2 mol% of 3-

hydroxyhexanoate was produced from glucose and octanoic acid, using bacterial fermentation,

as described by Xuan et al. [70,72]. The weight average molecular weight of the PHO,

determined by triple-detector size exclusion chromatography (SEC), was 172,000 Da with a

dispersity of 1.75 [126]. PLA (grade 3251D, MFI 35 g/10 min at 190 °C/ 2.16 kg) was obtained

from Natureworks®. Triallyl trimesate (TAM, 98%, Monomer Polymer Inc.), dicumyl peroxide

(DCP, 98%, Sigma-Aldrich), acetone (Sigma-Aldrich), and tetrahydrofuran (THF, Sigma-Aldrich)

were used as received.

6.2.2 Compounding

PLA and PHO were dried in a vacuum oven at 100 °C and at room temperature respectively, to

remove moisture. PLA/PHO blends containing 0-20 wt.% PHO were compounded in a DSM

microcompounder at 180 °C for 3 min at a screw speed of 100 rpm. The compounder was

operated under nitrogen blanket to limit polymer degradation. After compounding, the strands

were quenched in cold water before chopping into pellets. The neat materials were

compounded under the same conditions for comparison.

Page 89: Preparation and characterization of biopolymer compounds

71

Peroxide-degraded PHO and PLA/PHO blends were prepared by coating ground PHO and PLA

powders with an acetone solution containing DCP and allowing the solvent to evaporate. The

resulting mixtures of PHO were charged to a DSM micro-compounder, equipped with twin-co-

rotating screws, at 180 °C at 100 rpm for 6 min. Coagent-modified PHO and PLA reacted with

DCP and TAM were prepared as described above from PHO or PLA, yielding various

compositions containing 0.2-0.5 wt.% DCP and amounts of TAM ranging from 0.5-2 wt.%.

Similarly coagent-modified PLA/PHO blends containing DCP and TAM were prepared. The

various compounds are designated using the name of the polymer, followed by the amounts of

DCP and TAM (i.e. PHO/0.3/1 denotes PHO reacted with 0.3 wt.% DCP and 1 wt.% TAM).

The gel content of coagent modified PHO and PLA/PHO blend was measured by dissolving the

material in chloroform for 7 h. The polymer was sealed in stainless steel wire (120 mesh)

according to ASTM D 2765. The material was left to stand for 1 h and subsequently dried

overnight in a vacuum oven at room temperature. The % gel content was calculated using

equation (6.1).

100sample of weight Initial

sample of weight Finalcontent Gel (6.1)

6.2.3 Characterization

6.2.3.1 Rheology

Compression molded discs, 25 mm diameter and 2 mm thick, were prepared using a Carver

press. The linear viscoelastic properties were measured in the oscillatory mode using a stress

controlled rheometer (Visco Tech from Reologica). Frequency sweeps were conducted at 180 °C

using 20 mm parallel plates.

Page 90: Preparation and characterization of biopolymer compounds

72

Reactively modified PLA/PHO blends were further characterized in simple uniaxial extension

using an SER-2 universal testing platform from Xpansion Instruments hosted on the MCR-301

Anton Paar rheometer. Measurements were conducted at 180 °C at Hencky strain rates ranging

between 0.10 and 10 s-1. Specimens were prepared by compression molding the polymer

samples between polyester films to a gauge of about 0.75 mm, using a hydraulic press.

Individual polymer specimens were then cut to a width of 10 mm. Linear viscoelastic (LVE)

oscillatory measurements obtained at 180 °C were used to calculate the LVE stress growth

curve and check the consistency of the extensional measurements.

6.2.3.2 Differential scanning calorimetry (DSC)

DSC experiments were performed using a Q100 DSC from TA Instruments, under dry nitrogen.

Since MCL PHAs crystallize slowly, the samples were preconditioned to eliminate their thermal

history as follows: the polymer was heated at 100 °C for 10 min in a convection oven, and then

kept at room temperature for two weeks before characterization. Samples weighing 10-12 mg

were sealed in aluminum hermetic pans, equilibrated at -70 °C and kept isothermally for 5 min.

Afterwards they were heated to 200 °C at a rate of 5 °C/min and held isothermally for 3 min

before cooling to -70 °C at a rate of 5 °C/min. The samples were finally reheated to 200 °C at a

rate of 5 °C/min. The % crystallinity of the polymers, Xc, was estimated using equation (6.2).

100100

H

HHX ccm

c (6.2)

where, ΔHm is the enthalpy of fusion, ΔHcc is the exothermic enthalpy (cold crystallization)

recorded during DSC heating cycle and ΔH100 is the theoretical fusion enthalpy of a 100%

crystalline polymer. The heat of fusion for 100% crystalline PLA is 93.6 J/g [109].

Page 91: Preparation and characterization of biopolymer compounds

73

6.2.3.3 Heat deflection temperature (HDT)

Specimens (127 mm X 13 mm X 3 mm) were prepared by compression molding using a Carver

press under 5000 N force, at 200 °C with a residence time of 3 min, then quenched in cold

water. Specimens were lowered in a silicon oil bath and the temperature was raised from 23 oC

at a heating rate of 120 oC/h. until 0.25 mm deflection occurred under a load of 1.82 MPa, in

accordance with ASTM D 648. At least three specimens were tested and the average value was

reported.

6.2.3.4 Mechanical properties

Specimens for mechanical property characterization were prepared by compression molding

using a Carver press under 5000 N force, at 200 °C and a residence time of 3 min, then

quenched in cold water. All specimens were conditioned at room temperature for 48 h after

compression molding, prior to mechanical testing. Tensile tests were conducted in accordance

with ASTM D638 using standard type V test specimens, with an Instron 3369 Universal tester, at

a cross head speed (CHS) of 5 mm/min. The average of five measurements is reported. Un-

notched Izod impact tests were conducted in accordance with ISO 180 using standard

specimens on a SATEC Instron machine and the average of five specimens are reported.

6.2.3.5 Scanning electron microscopy

Blend morphologies were observed using a JEOL JSM-840 scanning electron microscope.

Samples were first hot-pressed at 200 oC for 3 min, then immersed in liquid nitrogen for 3 min

before brittle fracture. The MCL PHA phase was etched in acetone overnight at room

temperature. The coagent modified blend samples were examined without etching as MCL PHA

was partially cross-linked and could not be dissolved.

Page 92: Preparation and characterization of biopolymer compounds

74

6.2.3.6 Hot stage microscopy

Isothermal crystallization experiments were performed using a Linkam CSS 450 hot stage

mounted on an Olympus BX51 optical microscope. The sample was first heated to 200 °C at a

rate of 30 °C /min and held for 10 min to eliminate the heat history. The melt was then cooled

to 135 °C at 30 °C/min. The crystallization process was recorded isothermally at 135 °C using a

Sony ExwaveHAD 3 CCD digital recorder.

6.3 Results and Discussion

6.3.1 Blends of PLA with PHO

PLA and PHO form an immiscible blend system, having droplet-matrix morphology at

compositions up to 30 wt.% PHO, as shown in Figure 6.1. These blends have very coarse

morphology, with the average dispersed domain size changing from 1.5 ( 0.2) µm for 95/05

PLA/PHO blend to 5 ( 0.6), 7 ( 2) and 7.6 ( 2) µm for 90/10, 85/15 and 80/20 blend

respectively. This is consistent with the findings reported in Chapter 4 on PHB/PHO blends, and

is attributed to the significant viscosity mismatch between the blend components. PLA and PHO

had an almost Newtonian behavior, with viscosities of 660 Pa.s and 13 Pa.s for PLA and PHO

respectively, at the compounding temperature of 180 °C, resulting in a viscosity ratio (defined

as the ratio of the viscosity of the dispersed phase over the viscosity of the matrix) of 0.02.

Generally viscosity ratios as close as possible to 1 are required to achieve optimum blend

morphology, whereas viscosity ratios much higher or lower than 1 result in coarse

morphologies and tendency toward coalescence during melt compounding.

Page 93: Preparation and characterization of biopolymer compounds

75

Table 6.1 summarizes the mechanical properties of the blends. Significant improvement was

observed in the elongation and impact properties, accompanied by a decrease in Young’s

modulus and tensile strength. The decline in these properties was associated with the decrease

in crystallinity of the blend, which is expected because of the addition of an amorphous minor

phase.

Figure 6.1 Scanning electron microscope images of PLA blend containing a) 5 wt.%, b) 10 wt.%,

c) 15 wt.% and d) 20 wt.% of PHO.

From Table 6.1 it is obvious that the optimum levels of impact strength were obtained at

compositions between 10 and 15 wt.% PHO, whereas the values decreased when higher

Page 94: Preparation and characterization of biopolymer compounds

76

amounts were added. This is attributed to the morphology of the blend, which became much

coarser at compositions above 10 wt.% PHO, as shown in Figure 6.1.

Table 6.1 Mechanical properties of PLA and PLA/PHO blends

As mentioned in the previous chapters, low melt viscosities, viscosity mismatch and absence of

melt strength are factors that make processing of biopolyesters and their blends difficult,

resulting in a need for chain extension. In Chapter 4, we proposed peroxide-mediated cross-

linking of the PHO dispersed phase to achieve higher viscosity, and therefore to improve the

morphology of the blend. However, this approach produced high gel contents and limited

property improvements in blends with PHB. A similar attempt at blending peroxide cross-linked

PHO with PLA, shown in Appendix A, also resulted in limited success.

In the present chapter, a reactive modification procedure was implemented using peroxide and

coagent, similar to the approach described in Chapter 5, to achieve branching and therefore

PHO

( wt.%)

Tensile stress

(MPa)

Young's

Modulus

(MPa)

Elongation

at break (%)

Unnotched

Impact

(KJ/m2)

Crystallinity

(%)

0 74 (±3) 670 (±74) 14 (±1) 32 (±5) 24

5 56 (±4) 582 (±68) 24 (±10) 63 (±6) 17

10 50 (±4) 696 (±29) 35 (±15) 65 (±5) 16

15 45 (±3) 571 (±22) 47 (±10) 53 (±9) 16

20 34 (±3) 442 (±34) 24 (±3) 40 (±5) 16

Page 95: Preparation and characterization of biopolymer compounds

77

chain extension, while avoiding excessively high gel contents. The following sections describe

the effects of coagent modification on PHO, PLA and PLA/PHO blends.

6.3.2 Reactive modification of PHO

PHO was reacted with various amounts of TAM and DCP, aiming in general at choosing

formulations that would use the least amounts of reagents possible, while still achieving

acceptable improvements in viscosity and elasticity, without excessive gel contents.

As shown in Figure 6.2, the flow characteristics of PHO remained unaltered when reacted with

0.3 wt.% DCP (PHO/0.3). The absence of changes in the presence of DCP is opposite to what

was observed in Chapter 4, when lauroyl peroxide was used. This suggests that in the presence

of DCP chain extension/cross-linking in the presence of peroxide is counteracted by significant

chain scission, possibly because of the higher compounding temperatures needed for DCP

compared to lauroyl peroxide.

When the multifunctional coagent, TAM, was added to the formulation, significant changes

were noted in the rheological properties of PHO. Two distinct groupings are noted in Figure

6.2a. When TAM amounts up to 0.75 wt.% were added (sample PHO/0.3/0.75), the increase in

complex viscosity was relatively small. On the contrary, addition of 1 wt.% TAM resulted in a

pronounced increase in viscosity, shear thinning behavior, and elasticity. Above this level of

TAM, the properties showed a tendency to plateau.

Page 96: Preparation and characterization of biopolymer compounds

78

Figure 6.2 Effect of TAM content on the rheological properties of PHO with DCP content

remaining constant a) complex viscosity b) storage modulus and c) tan δ

1 10 100

Co

mp

lex

Vis

co

sit

y (

Pa

s)

104

103

102

101

100

1 10 100

Sto

rag

e M

od

ulu

s (P

a)

104

102

103

101

10-2

10-1

100

1 10 100

tan

δ

Frequency (rad/s)

PHO PHO/0.3PHO/0.3/0.5 PHO/0.3/0.75PHO/0.3/1 PHO/0.3/2

103

102

100

10-1

101

b)

a)

c)

Page 97: Preparation and characterization of biopolymer compounds

79

Reactive modification was accompanied by changes in the appearance of the extrudate. The

PHO formulations containing low amounts of TAM were sticky and did not form strands when

extruded. On the contrary, addition of 1 wt.% TAM resulted in the formation of strands that

were not sticky and were easy to handle for further processing and characterization as shown in

Figure 6.3.

Figure 6.3 a) unmodified PHO after extrusion b) PHO/0.3/1 after extrusion

In an effort to achieve an optimum formulation, the amount of DCP was varied, while keeping

the amount of TAM constant at 1 wt.% (Figure 6.4). From this figure, it is obvious that amounts

of DCP above 0.3 wt.% were necessary to obtain significant improvements in complex viscosity

and elasticity. However the material reacted with 0.5 wt.% DCP was highly cross-linked, with a

gel content of 42 %. Based on the above results, the formulation containing 0.3 wt.% of DCP

with 1 wt.% of TAM showed the best improvement in viscosity, while maintaining a moderate

gel content of 23%.

a) b)

Page 98: Preparation and characterization of biopolymer compounds

80

Figure 6.4 Effect of DCP amount on a) complex viscosity b) storage modulus and c) tan δ of

coagent modified PHO (PHO 0.3/1 and PHO 0.5/1 yielded 23 and 42 % gel respectively)

1 10 100

Co

mp

lex V

isco

sit

y (

Pa s

)

Frequency (rad/s)

PHO/0.2/1

PHO/0.3/1

PHO/0.5/1

102

1 10 100

Sto

rag

e M

od

ulu

s (P

a)

Frequency (rad/s)

10-2

105

104

103

102

101

100

10-1

1 10 100

tan

δ

Frequency (rad/s)

103

105

104

103

101

100

102

101

100

10-1

10-2

a)

b)

c)

Page 99: Preparation and characterization of biopolymer compounds

81

In spite of the dramatic changes in the rheological properties, the thermal properties of the

reactively modified PHO, including the glass transition temperature (Tg) remained unaltered,

whereas the material remained highly amorphous, with no obvious melting transition.

6.3.3 Reactive modification of PLA

The reactive modification of the PLA matrix to provide branched PLA has been described in

detail in Chapter 5. Figure 6.5 shows the rheological properties of PLA reacted with two

different amounts of coagent and comparison with PHO/0.3/1. The formulation reacted with 2

wt.% TAM had very high viscosity, and appeared fully crosslinked and hard to process. On the

contrary, PLA/0.3/1 had negligible gel content, therefore this composition was deemed suitable

for the blend compositions described in Section 6.3.3. below.

Figure 6.5 Effect of coagent modification on the complex viscosity of PLA and PHO

1 10 100

Co

mp

lex V

isco

sit

y

(Pa s

)

Frequency (rad/s)

PLA PLA/0.3/1

PLA/0.3/2 PHO/0.3/1

102

102

104

103

101

104

102102

102

102

104

103

101

104

101

Page 100: Preparation and characterization of biopolymer compounds

82

The extensional properties of the PLA/0.3/1 are summarized in Figure 6.6. As explained in

Chapter 5, reactive modification resulted in substantial strain hardening, which provides solid

evidence of a branched structure.

Figure 6.6 Tensile stress growth coefficient (ηE+) of PLA/0.3/1 and (PLA/PHO)/0.3/1 as a

function of strain rate and time at Hencky strain rates of 0.1, 1 and 10 s-1 at 180 °C. Curves are

shifted by an arbitrary factor for the sake of clarity. Dotted lines represent the LVE envelop for

each sample.

In addition to the enhancements in viscosity and strain hardening, the coagent modified PLA

had substantially different thermal properties compared to the neat PLA, as shown in Table 6.2,

including the appearance of a sharp crystallization peak in the DSC exotherm (Figure 6.7),

disappearance of the cold crystallization peak and increase in crystallinity.

0.01 0.1 1 10 100

ȠE

+(P

a s

)

t (s)

0.1 s-11 s-110 s-1

PLA/PHO/0.3/1

105

106

103

104

PLA/0.3/1

10x

102

Page 101: Preparation and characterization of biopolymer compounds

83

Table 6.2 Thermal properties of neat, DCP and coagent modified PHO, PLA and PHO/PLA blend

TM (°C) TC (°C) TCC (°C) Crystallinity (%)

PLA 173 NA 109 24

PLA/0.3 170 NA 94 35

PLA/0.3/1 169 133 NA 52

PLA/PHO 170 NA 105 16

(PLA/PHO) /0.3 170 NA 104 17

(PLA/PHO) /0.3/1 169 138 NA 56

As reported in Chapter 5, the crystallization kinetics of PLA appeared significantly enhanced,

both in isothermal, and non-isothermal experiments. This finding was attributed to the

formation of a separate phase of coagent-rich particles that forms upon reactive modification,

which results in a nucleating effect [55].

Page 102: Preparation and characterization of biopolymer compounds

84

Figure 6.7 DSC (a) cooling exotherm (b) heating endotherm of coagent-modified PLA and

PLA/PHO blends

80 100 120 140 160 180

He

at

Flo

w A

. U

.

Temperature ( C)

PLA/PHO

PLA/PHO/0.3

PLA/PHO/0.3/1

PLA/0.3/1

80 100 120 140 160 180

He

at

Flo

w A

.U.

a)

b)

Page 103: Preparation and characterization of biopolymer compounds

85

Further evidence of the altered crystalline structure of these materials upon coagent

modification is provided by the optical microscope images in Figure 6.8, which depict PLA and

PLA/0.3/1 samples crystallized under isothermal conditions at 135 °C. The unmodified PLA did

not show evidence of crystal structure formation when cooled for 5 min, whereas the coagent-

modified sample is characterized by the formation of a dense spherulitic structure.

Figure 6.8 Hot stage microscopy of a) PLA, b) PLA/0.3/1 at 135 °C

The profound changes in the thermal and rheological properties of PLA are expected to have a

major impact on the properties of reactively compounded blends of PLA with PHO, when PLA is

the matrix phase. These blends are presented in section 6.3.4 – 6.3.7 below.

6.3.4 Reactive compounding of PLA with PHO

Based on the analysis presented in section 6.3.1, blends of PLA/PHO containing 10 wt.% PHO,

were used as this was identified as the optimum composition in terms of mechanical

properties. An additional reason for keeping the PHO content as low as possible is its high cost.

For simplicity these blends will be noted below as PLA/PHO.

a) b)

Page 104: Preparation and characterization of biopolymer compounds

86

6.3.5 Thermal and rheological properties

Reaction of PLA/PHO blends with various amounts of peroxide and coagent resulted in similar

trends as noted in the sections above, i.e. increase in the complex viscosity, loss of Newtonian

plateau, increased shear thinning and elasticity of the blends. The values of all viscoelastic

properties were in-line with those seen for the PLA matrix at equivalent contents (Figures 6.5

and 6.9), with what appears to be limited influence of the minor PHO dispersed phase.

Similarly, the extensional stress growth data showed strain hardening behavior, following the

same trends as those seen for the branched PLA matrix (Figure 6.6).

In agreement with the findings reported for the reactively modified PLA, the modified PLA/PHO

showed a clear and sharp crystallization peak at 138 °C (Figure 6.7a). This was accompanied by

the disappearance of the cold crystallization peak, and significantly higher crystallinities

compared to the unreacted materials (Figure 6.7b). The thermal properties of all the reactively

modified compounds and their unreacted counterparts are summarized in Table 6.2.

6.3.6 Blend morphology

Compounding of PLA and PHO in the presence of DCP alone, resulted in a significant reduction

in the PHO domain size, as seen in Figure 6.10. The size of the PHO dispersed domains

decreased from 5 ( 0.6) µm to 1 ( 0.2) and 1.2 ( 0.2) µm when 0.5 and 1 wt. % of DCP was

added. Since, based on the rheological evaluations presented above, DCP alone did not

influence the rheological properties of the constituents of the blend, the most likely

explanation for this observation is a compatibilizing effect in the presence of DCP. This could be

attributed to the formation of small amounts of copolymer, by cross-termination of free radical

chains of PLA and PHO during the compounding procedure. In-situ compatibilization by reactive

Page 105: Preparation and characterization of biopolymer compounds

87

blending of biopolyesters through reaction with peroxides to produce blends with finer

morphologies has been reported previously [17,123-125].

Figure 6.9 Effect of DCP and TAM on a) complex viscosity b) storage modulus and c) tan δ of

PLA/PHO blends

1 10 100

Co

mp

lex

Vis

co

sit

y (

Pa

s)

Frequency (rad/s)

104

103

102

1 10 100

Sto

rag

e M

od

ulu

s (P

a)

Frequency (rad/s)

106

105

104

102

103

101

1 10 100

tan

δ

Frequency (rad/s)

PLA/PHO PLA/PHO/0.3

PLA/PHO/0.3/0.3 PLA/PHO/0.3/1

PLA/PHO/0.5/1

102

101

100

10-1

b)

a)

c)

Page 106: Preparation and characterization of biopolymer compounds

88

Ma et al. [124,127] and Dong et al. [125] attributed the compatibilization effect to the

combination of macroradicals that form in the presence of DCP via hydrogen abstraction. The

macroradicals may further recombine to form complex products, including copolymer at the

interface, resulting in a compatibilization effect.

Addition of TAM to the formulation resulted in further refinement in the morphology, which is

may be attributed to the improved viscosity of the PHO. The PHO domain size reduced from 1.5

( 0.2) to 0.55 ( 0.08) µm and from 5 ( 0.6) to 1 ( 0.16) µm for 95/5 and 90/10 PLA/PHO

blends respectively. In case of 80/20 the blend morphology changed from droplet-matrix to co-

continuous. Figure 6.11 shows the altered morphologies obtained in the presence of TAM, in

blends containing various amounts of PHO. In addition to the finer morphology noted

previously, these blends exhibited a different, co-continuous morphology at a PHO content of

20 wt.%. Such changes in morphology are common in thermoplastic vulcanizates, consisting of

PP and EPDM [128] but have never been noted for these biopolyester blends.

Figure 6.10 Scanning electron microscopy of PLA/PHO (90/10) blend a) unmodified b)

(PLA/PHO)/0.5 c) (PLA/PHO)/1

20mm 20mm20mm

a) b) c)

Page 107: Preparation and characterization of biopolymer compounds

89

Figure 6.11 Effect of coagent modification on morphology of PLA/PHO blends a) 95/05 b) 90/10

c) 80/20 (wt./wt.%) (samples reacted with coagent were not etched); Top row without coagent;

bottom row with coagent

6.3.7 Mechanical properties

The reactively modified PLA/PHO blend had better tensile strain and unnotched impact

strength compared to pristine PLA, however, its properties were not as good as the unmodified

blend, in spite of the improved morphology (Table 6.3). The gel content of coagent modified

PLA/PHO blend was 20 %, revealing the presence of cross-linked chains in the modified blend.

These gels might have compromised the ductility of the blend, thus explaining the drop in

ductility and impact strength compared to the unmodified blend. Addition of PHO to PLA, as

well as reactive modification, did not affect its heat deflection temperature (HDT) (Table 6.3).

a) b) c)

Page 108: Preparation and characterization of biopolymer compounds

90

Table 6.3 Mechanical properties and heat deflection temperature of neat and coagent modified

PLA, and PLA/PHO blends

These findings suggest that avoiding gels is crucial for the optimization of the properties of

these materials. Cross-linking/gelation is also expected to affect negatively the biodegradability

of these materials. Therefore optimization of the compositions to avoid the formation of gels

should be a high priority.

These results have shown that coagent modification improves significantly the processability of

the materials, by improving the melt strength and crystallization rates, while the mechanical

properties remain relatively unaffected when the specimens are prepared under the same

conditions. The differences noted in thermal properties and morphology however suggests that

processing conditions during the solidification stage may be tuned to further impart change in

the mechanical properties. This should be a topic of further investigation.

Tensile

stress

(MPa)

Tensile

strain

(%)

Unnotched

Impact

(KJ/m2)

Young's

Modulus

(MPa)

HDT

(°C)

PLA 74 (±3) 14 (±1) 32 (±5) 670 (±74) 55

PLA/0.3/1 77 (±1) 13 (±2) 35 (±2) 837 (±74) 56

PLA/PHO 50 (±1) 35 (±15) 65 (±5) 696 (±7) 55

(PLA/PHO)/0.3/1 49 (±3) 24 (±4) 55 (±3) 677 (±34) 54

Page 109: Preparation and characterization of biopolymer compounds

91

6.4 Conclusions

Addition of PHO to PLA increased the impact strength and elongation at break of PLA, at the

expense of the Young’s modulus, while the HDT remained unaffected. The droplet-matrix

morphology of the blends was coarse, because of the very low viscosity of PHO, resulting in a

viscosity mismatch between the blend components.

The viscosity of PHO was successfully increased through partial cross-linking, using solvent-free

chemical modification using DCP and TAM coagent. In addition to the substantial increase in

viscosity and melt elasticity, the resulting product exhibited improved extrudate appearance.

Reactively modified PLA/PHO in the presence of DCP and TAM displayed the enhancements in

strain hardening and crystallization rates, previously observed for the matrix modified PLA.

Furthermore these blends had finer morphology, which was attributed to a compatibilizing

effect possibly arising from copolymer formation at the interface. Coagent modification further

resulted in changes in the morphology, and possible phase inversion in blends containing higher

PHO contents.

Page 110: Preparation and characterization of biopolymer compounds

92

Chapter 7 Thesis overview

7.1 Thesis overview

Ongoing need for alternate options to conventional petroleum based polymers has resulted in

significant attention to various biopolyesters, including poly(lactic acid) (PLA) and poly-(3-

hydroxyalkanoates) (PHAs), which are bioderived and biodegradable. However, most of these

materials typically suffer from high production costs, brittleness, slow crystallization rates and

low melt strength, which restrict their processability under common polymer processing

operations. This thesis focused on the improvement of properties and processability of

biopolyesters through blending and reactive compounding. Specifically the potential of

elastomeric medium-chain-length (MCL) PHAs, as potential impact modifiers for PLA and brittle

poly-3-hydroxybutyrate (PHB) was assessed, using conventional melt compounding.

In spite of the relatively high molar masses (ranging from 18,200 to 172,000 g/mole), MCL PHAs

melts have low viscosity, presumably due to their helical conformation. Melt blending of these

materials with PLA and PHB resulted in coarse blend morphologies due to the large viscosity

mismatch, however, remarkable improvement was observed in ductility and impact strength of

PHB and PLA. As expected these improvements were accompanied by a decline in tensile stress

and Young’s modulus. Free-radical mediated cross-linking of polyhydroxyoctanoate (PHO) using

a peroxide resulted in improved blend morphology, because of the increased viscosity of PHO;

however the impact properties did not show further improvements, presumably because of the

high gel content. Furthermore chain extension using epoxidized PHO (ePHO) were explored.

This approach lowered the interfacial tension between PLA and PHO, because of the

Page 111: Preparation and characterization of biopolymer compounds

93

interactions between the epoxy groups of ePHO with the carboxylic acid end group of PLA, thus

resulting in improved compatibility. However the morphology remained coarse, because of the

extremely low viscosity of ePHO.

The inability of the previously suggested approaches to produce substantial improvements in

the processability and properties of these polyesters led to the development of a simple

reactive compounding approach, involving reaction with a peroxide and coagent. This method

resulted in substantially improved crystallization rates, viscosity, elasticity and melt strength of

PLA and its blends with PHO, as well as improved the compatibility between the blend

components, resulting in a very fine morphology. This approach has resulted in blends having

simultaneously improved toughness, increased viscosity, strain hardening and crystallinity and

represents a significant technological advance in these materials.

These results suggest that simple free-radical mediated reactive compounding of biopolyesters

in the melt state can produce the rheological enhancements needed in processes such as film

blowing and casting, blow molding, thermoforming and foaming, as well as the enhanced

crystallization rates needed in injection molding, thus significantly broadening the applicability

of these polymers in conventional polymer processing operations. This should enable the

introduction of these compounds to new products/markets and may lead to the possible

identification of new applications for these biopolymers.

7.2 Conclusions

Absolute molecular weight (MW) distributions were determined for different MCL PHAs with

predominantly 3-hydroxyoctanoate (PHO), 3-hydroxynonanoate (PHN) or 3-

hydroxydodecanoate (PHDD) content via triple-detector size exclusion chromatography (SEC),

Page 112: Preparation and characterization of biopolymer compounds

94

combined with analyses using various detectors, using tetrahydrofuran (THF) as the carrier

solvent. Unlike with the short-chain-length (SCL) PHB, the uncorrected polystyrene calibration

in THF provided a good estimate (within 10 %) of absolute MW values for the tested MCL PHAs,

irrespective of side chain length. Weight-average MW values ranged from 172,000 Da for PHO

to 18,200 for PHN with 30 mol% 3-hydroxyheptanoate, and dispersities of all samples were

close to two. Melt viscosity data suggested an entanglement molecular weight around 8 X 104

Da, significantly higher than most thermoplastics.

Blends of PHO with PHB, were prepared by melt compounding. Coarsening of the droplet-

matrix morphology of the blends was noted as the PHO content increased beyond 5 wt.%; this

was attributed to the significant viscosity mismatch between the components. Addition of PHO

improved the thermal stability of the blends, reduced their crystallinity and resulted in shifts in

their melting and crystallization temperatures. The blends had improved tensile strain at break.

The unnotched impact strength showed a threefold increase at 30 wt.% PHO content. Cross-

linking of PHO using a peroxide initiator increased its viscosity, thus improving the morphology

and mechanical properties of the blends.

Blends of PLA with PHO were also investigated. In spite of the viscosity mismatch between the

blend components, addition of PHO resulted in increased elongation at break and impact

strength of PLA accompanying decrease in tensile strength and modulus. These improvements

can be explained by a reduction in the crystallinity of the PLA, whereas the melting point and

glass transition temperature (Tg) remained unaffected. Further increase of the PHO content

resulted in a drop in properties, and attributed to the coarse morphology. The heat deflection

temperature (HDT) of PLA did not change upon addition of 10 wt.% PHO. Peroxide cross-linking

Page 113: Preparation and characterization of biopolymer compounds

95

of PHO to increase its melt viscosity, and introduction of epoxy groups led to only moderate

improvements in blend properties.

In order to improve melt and crystallization properties, PLA was chemically modified by radical

mediated solvent-free, peroxide-initiated grafting of triallyl trimesate (TAM) coagent in the

melt state. When compared with the parent material and with PLA samples treated with

peroxide alone, coagent-modified materials demonstrated higher molar mass and improved

melt rheological properties, including substantial improvements in melt elasticity and strain

hardening under uniaxial extension. The properties of coagent modified PLA were compared to

those PLA modified by a multi-functional epoxide oligomeric chain extender (Joncryl®).

Although the rheological properties were comparable, the coagent-modified material

demonstrated significantly enhanced crystallinity and crystallization rates. The appearance of a

distinct crystallization exothermic peak and the disappearance of the cold crystallization

temperature point to a nucleation effect in the coagent modified PLA, which together with the

rheological enhancements can promote the processability of this material in conventional

thermoplastics operations.

Reactive compounding in the presence of the dicumyl peroxide (DCP) and TAM was also

evaluated in PHO and its blends with PLA. The viscosity and elasticity of PHO increased

substantially following reactive compounding indicative of partial cross-linking, while it retained

its amorphous nature. In addition to the substantial increase in viscosity and melt elasticity, the

resulting product exhibited improved extrudate appearance. Reactively modified PLA/PHO

blends in the presence of DCP and TAM displayed enhancement in strain hardening and

crystallization rates, previously observed for the matrix, modified PLA. Furthermore these

Page 114: Preparation and characterization of biopolymer compounds

96

blends had finer morphology, which was attributed to a compatibilizing effect possibly arising

from copolymer formation at the interface. Coagent modification further resulted in changes in

the morphology, and possible phase inversion in blends containing higher PHO content, while

the mechanical properties and HDT remained relatively unaffected by this coagent

modification.

7.3 Significant contributions

This thesis has completed a very thorough and complete characterization of various MCL PHAs,

which are possible candidates to replace conventional elastomeric polymers. For the very first

time, the true molecular weight of MCL PHAs was determined. Based on the findings, molar

mass determinations based on polystyrene standards can be used as the molecular weight of

these materials.

MCL PHAs have very low viscosities and crystallize very slowly, requiring 8-12 h to solidify after

processing. This makes their handling and post-processing very difficult. The work in this thesis

developed a method to improve processing and handling of MCL PHA, through a simple

reactive modification using a peroxide and coagent.

It was shown that PHO, a bioderived and biodegradable polymer, can be used to impact modify

brittle biopolymers to offer a complete biosystem. Addition of PHO improved elongation and

impact properties of brittle PLA and PHB.

The most important contribution of this thesis is that it provided a solution to improve the melt

strength of PLA and its blends with PHO. This should extend its use in applications such as blow

molding, thermoforming, blown film and foaming that involve stretching. The improvements in

the total crystallinity and crystallization rates of PLA and its blends with PHO without the need

Page 115: Preparation and characterization of biopolymer compounds

97

to add nucleating agents, are another significant finding of this thesis, which will likely have

broad implications in their potential to extend to new applications.

7.4 Recommendation for future work

i) Optimization of coagent content is needed to reduce gel content so that the

elastomeric properties of MCL PHA can be retained and its ability to impact modify

brittle polymers can fully be utilized.

ii) Most of the MCL PHAs, with the exception of poly(3-hydroxydodecanoate) (PHDD)

were highly amorphous. It may be worthwhile to pursue coagent modification of

PHDD, to investigate the potential to impart some improved crystallinity to this

material.

iii) The solid-state properties of coagent modified MCL PHAs should be investigated, as

these materials could be used as replacements for conventional thermoplastic

elastomers.

iv) The effect of coagent modification on pristine PHB and its blends with MCL PHA can

be investigated in detail to see whether the properties of these blends can be

enhanced.

v) Further detailed investigations of the improvements in crystallization in the

presence of coagent should be carried out, including detailed optical microscopy, X-

ray diffraction (XRD) analysis, as well as investigation of coagent-rich particle

formation in these compounds. The improvements in the morphology of PLA/MCL

PHA blends in the presence of peroxide and coagent should also be studied.

Page 116: Preparation and characterization of biopolymer compounds

98

vi) The mechanical properties of coagent modified materials should be influenced by

the differences in their crystallinity. Detailed investigation of this, including different

cooling rates during sample preparation, as well as different molding procedures,

(injection and compression molding) should be carried out to exploit this effect.

Page 117: Preparation and characterization of biopolymer compounds

99

References

1. Azapagic A, Emsley A, Hamerton I. Polymer – The environment and sustainable development.

Hamerton I, Wiley, 2003.

2. Gerald S. Degradable polymers: principles and applications. Netherlands: Kluwer academic

publishers, 2002.

3. Opportunities in chemistry by National research council. Washington DC: National academy

press, 1985.

4. Lee J, McCarthy S. Biodegradable poly(lactic acid) blends with chemically modified

polyhydroxyoctanoate through chain extension. J Polym Environ 2009;17(4):240-247.

5. Lim L, Auras R, Rubino M. Processing technologies for poly(lactic acid). Progress in Polymer

Science 2008;33(8):820-852.

6. Hanggi U. Requirements on bacterial polyesters as future substitute for conventional plastics

for consumer goods. FEMS Microbiol Rev 1995;16(2-3):213-220.

7. Lemoigne M. Ann.Inst.Pasteur,Paris. 1925,(39):144.

8. Steinbuchel A, Valentin H. Diversity of bacterial polyhydroxyalkanoic acids. FEMS Microbiol

Lett 1995;128(3):219-228.

9. Takagi Y, Yasuda R, Yamaoka M, Yamane T. Morphologies and mechanical properties of

polylactide blends with medium chain length poly(3-hydroxyalkanoate) and chemically

modified poly(3-hydroxyalkanoate). J Appl Polym Sci 2004;93(5):2363-2369.

10. Gassner F, Owen A. Physical properties of poly(beta-hydroxybutyrate) poly(epsilon-

caprolactone) blends. Polymer 1994;35(10):2233-2236.

Page 118: Preparation and characterization of biopolymer compounds

100

11. Grassie N, Murray EJ, Holmes PA. The thermal degradation of poly(-(d)-β-hydroxybutyric

acid): Part 1—Identification and quantitative analysis of products. Polym Degrad Stab

1984;6(1):47-61.

12. Hong S, Gau T, Huang S. Enhancement of the crystallization and thermal stability of

polyhydroxybutyrate by polymeric additives. J Therm Anal Calorim 2011;103(3):967-975.

13. Huisman G, Deleeuw O, Eggink G, Witholt B. Synthesis of poly-3-hydroxyalkanoates is a

common feature of fluorescent pseudomonads. Appl Environ Microbiol 1989;55(8):1949-1954.

14. Jiang X. Process development for the production and separation of medium-chain-length

poly ( 3-hydroxyalkanoates) by Pseudomonas putida KT2440. Ph.D. Thesis, 2010.

15. Diard S, Carlier J, Ageron E, Grimont P, Langlois V, Guerin P, Bouvet O. Accumulation of

poly(3-hydroxybutyrate) from octanoate, in different Pseudomonas belonging to the rRNA

homology group I. Syst Appl Microbiol 2002;25(2):183-188.

16. Sun Z, Ramsay JA, Guay M, Ramsay BA. Fermentation process development for the

production of medium-chain-length poly-3-hyroxyalkanoates. Appl Microbiol Biotechnol

2007;75(3):475-485.

17. Gagnon K, Lenz R, Farris R, Fuller R. Crystallization behavior and its Influence on the

mechanical properties of a thermoplastic elastomer produced by Pseudomonasoleovorans.

Macromolecules 1992;25(14):3723-3728.

18. Utracki LA. Polymer blends handbook. Kluwer academic publishers, 2002.

19. Parulekar Y, Mohanty A. Biodegradable toughened polymers from renewable resources:

blends of polyhydroxybutyrate with epoxidized natural rubber and maleated polybutadiene.

Green Chem 2006;8(2):206-213.

Page 119: Preparation and characterization of biopolymer compounds

101

20. Avella M, Martuscelli E. Poly-d-(-)(3-hydroxybutyrate) poly(ethylene oxide) blends - phase-

diagram, thermal and crystallization behavior. Polymer 1988;29(10):1731-1737.

21. Hori Y, Takahashi Y, Hongo H, Yamaguchi A, Hagiwara T. Takasago Int. 1996(EP 723 983).

22. Kumagai Y, Doi Y. Journal of Environmental Polymer Degradation 1993;1(2):81.

23. de Carvalho FP, Quental AC, Felisberti MI. Polyhydroxybutyrate/acrylonitrile-g-(ethylene-co-

propylene-co-diene)-g-s tyrene blends: Their morphology and thermal and mechanical

behavior. J Appl Polym Sci 2008;110(2):880-889.

24. Asrar J, Hill J. Biosynthetic processes for linear polymers. J Appl Polym Sci 2002;83(3):457-

483.

25. Ceccorulli G, Pizzoli M, Scandola M. Plasticization of bacterial poly(3-hydroxybutyrate).

Macromolecules 1992;25(12):3304-3306.

26. Riande E, Markovitz H, Plazek D, Raghupathi N. Viscoelastic behavior of polystyrene -

tricresyl phosphate solutions. Journal of Polymer Science Part C-Polymer Symposium

1975(50):405-430.

27. Ishikawa K, Kawaguchi Y, Doi Y. Plasticization of bacterial polyester by the addition of

acylglycerols and its enzymatic degradability. Kobunshi Ronbunshu 1991;48(4):221-226.

28. Withey R, Hay J. The effect of seeding on the crystallisation of poly(hydroxybutyrate), and

co-poly(hydroxybutyrate-co-valerate). Polymer 1999;40(18):5147-5152.

29. Alata H, Hexig B, Inoue Y. Effect of poly(vinyl alcohol) fine particles as a novel biodegradable

nucleating agent on the crystallization of poly(3-hydroxybutyrate). J Polym Sci Pt B-Polym Phys

2006;44(13):1813-1820.

Page 120: Preparation and characterization of biopolymer compounds

102

30. Maiti P, Batt CA, Giannelis EP. New biodegradable polyhydroxybutyrate/layered silicate

nanocomposites. Biomacromolecules 2007;8(11):3393-3400.

31. Kosan B, Michels C, Meister F. Dissolution and forming of cellulose with ionic liquids.

Cellulose 2008;15(1):59-66.

32. Ljungberg N, Cavaille JY, Heux L. Nanocomposites of isotactic polypropylene reinforced with

rod-like cellulose whiskers. Polymer 2006;47(18):6285-6292.

33. Danyadi L, Janecska T, Szabo Z, Nagy G, Moczo J, Pukanszky B. Wood flour filled PP

composites: Compatibilization and adhesion. Composites Sci Technol 2007;67(13):2838-2846.

34. Nanocrystalline cellulose as strong as Kevlar and extracted from plants is heading

to commercialization in Canada. http://nextbigfuture.com 2011.

35. Lucia LA. The nanoscience and technology of renewable biomaterials. John Wiley and Sons,

2009.

36. Rasal RM, Janorkar AV, Hirt DE. Poly(lactic acid) modifications. Prog Polym Sci

2010;35(3):338-356.

37. Saeidlou S, Huneault MA, Li H, Park CB. Poly(lactic acid) crystallization. Prog Polym Sci

2012;37(12):1657-1677.

38. Nampoothiri KM, Nair NR, John RP. An overview of the recent developments in polylactide

(PLA) research. Bioresour Technol 2010;101(22):8493-8501.

39. Zaman HU, Song JC, Park L, Kang I, Park S, Kwak G, Park B, Yoon K. Poly(lactic acid) blends

with desired end-use properties by addition of thermoplastic polyester elastomer and MDI.

Polym Bull 2011;67(1):187-198.

Page 121: Preparation and characterization of biopolymer compounds

103

40. Gui Z, Xu Y, Gao Y, Lu C, Cheng S. Novel polyethylene glycol-based polyester-toughened

polylactide. Mater Lett 2012;71:63-65.

41. Murariu M, Ferreira ADS, Pluta M, Bonnaud L, Alexandre M, Dubois P. Polylactide (PLA)-

CaSO4 composites toughened with low molecular weight and polymeric ester-like plasticizers

and related performances. Eur Polym J 2008;44(11):3842-3852.

42. Li H, Huneault MA. Effect of nucleation and plasticization on the crystallization of poly(lactic

acid). Polymer 2007;48(23):6855-6866.

43. Battegazzore D, Bocchini S, Frache A. Crystallization kinetics of poly(lactic acid)-talc

composites. Express Polym Lett 2011;5(10):849-858.

44. Yang G, Su J, Su R, Zhang Q, Fu Q, Na B. Toughening of poly(l-lactic acid) by annealing: The

effect of crystal morphologies and modifications. J Macromol Sci Part B-Phys 2012;51(1-3):184-

196.

45. Corre Y, Duchet J, Reignier J, Maazouz A. Melt strengthening of poly (lactic acid) through

reactive extrusion with epoxy-functionalized chains. Rheol Acta 2011;50(7-8):613-629.

46. Deenadayalan E, Lele AK, Balasubramanian M. Reactive extrusion of poly(l-lactic acid) with

glycidol. J Appl Polym Sci 2009;112(3):1391-1398.

47. Najafi N, Heuzey MC, Carreau PJ, Wood-Adams PM. Control of thermal degradation of

polylactide (PLA)-clay nanocomposites using chain extenders. Polym Degrad Stab

2012;97(4):554-565.

48. Villalobos M, Awojulu A, Greeley T, Deeter G. Oligomeric chain extenders for economic

reprocessing and recycling of condensation plastics, 2004.

Page 122: Preparation and characterization of biopolymer compounds

104

49. Bikiaris DN, Karayannidis GP. Chain extension of polyesters PET and PBT with two new

diimidodiepoxides .2. Journal of Polymer Science Part A-Polymer Chemistry 1996;34(7):1337-

1342.

50. Villalobos M, Awojulu A, Greeley T, Turco G, Deeter G. Oligomeric chain extenders for

economic reprocessing and recycling of condensation plastics. Energy 2006;31(15):3227-3234.

51. Al-Itry R, Lamnawar K, Maazouz A. Improvement of thermal stability, rheological and

mechanical properties of PLA, PBAT and their blends by reactive extrusion with functionalized

epoxy. Polym Degrad Stab 2012;97(10):1898-1914.

52. Dean KM, Petinakis E, Meure S, Yu L, Chryss A. Melt strength and rheological properties of

biodegradable poly(lactic acid) modified via alkyl radical based reactive extrusion processes. J

Polym Environ 2012;20(3):741-747.

53. Rytlewski P, Zenkiewicz M, Malinowski R. Influence of dicumyl peroxide content on thermal

and mechanical properties of polylactide. International Polymer Processing 2011;26(5):580-586.

54. Carlson D, Dubois P, Nie L, Narayan R. Free radical branching of polylactide by reactive

extrusion. Polym Eng Sci 1998;38(2):311-321.

55. Zhang Y, Tiwary P, Parent JS, Kontopoulou M, Park CB. Crystallization and foaming of

coagent-modified polypropylene: Nucleation effects of cross-linked nanoparticles. Polymer

2013;54(18):4814-4819.

56. Parent JS, Sengupta SS, Kaufman M, Chaudhary BI. Coagent-induced transformations of

polypropylene microstructure: Evolution of bimodal architectures and cross-linked nano-

particles. Polymer 2008;49(18):3884-3891.

Page 123: Preparation and characterization of biopolymer compounds

105

57. Parent JS, Bodsworth A, Sengupta SS, Kontopoulou M, Chaudhary BI, Poche D, Cousteaux S.

Structure-rheology relationships of long-chain branched polypropylene: Comparative analysis of

acrylic and allylic coagent chemistry. Polymer 2009;50(1):85-94.

58. Vaidya A, Pandey R, Mudliar S, Kumar M, Chakrabarti T, Devotta S. Production and recovery

of lactic acid for polylactide - an overview. Crit Rev Environ Sci Technol 2005;35(5):429-467.

59. Global biopolymers market analysis and forecasts to 2015. 2009;GMDCH0018MR.

60. Rai R, Yunos DM, Boccaccini AR, Knowles JC, Barker IA, Howdle SM, Tredwell GD, Keshavarz

T, Roy I. Poly-3-hydroxyoctanoate P(3HO), a medium chain length polyhydroxyalkanoate

homopolymer from Pseudomonas mendocina. Biomacromolecules 2011;12(6):2126-2136.

61. Nerkar M, Ramsay J, Ramsay B, Kontopoulou M. Morphology and mechanical properties of

blends of short-chain-length and medium-chain-length poly-3-hydroxyalkanoates. Polymer

Processing Society, 2012;358-359.

62. Akita S, Einaga Y, Miyaki Y, Fuita H. Solution Properties of poly(d-beta-hydroxybutyrate) .1.

biosynthesis and characterization. Macromolecules 1976;9(5):774-780.

63. Marchessault R, Okamura K, SU C. Physical properties of poly(beta-hydroxy butyrate) .2.

conformational aspects in solution. Macromolecules 1970;3(6):735-&.

64. Miyaki Y, Einaga Y, Hirosy T, Fujita H. Solution properties of poly(d-beta-hydroxybutyrate)

.2. light-scattering and viscosity in trifluoroethanol and behavior of highly expanded polymer

coils. Macromolecules 1977;10(6):1356-1364.

65. Cornibert J, Marchessault RH, Benoit H, Weill G. Physical properties of poly(beta-hydroxy

butyrate) .3. folding of helical segments in 2,2,2-trifluoroethanol. Macromolecules

1970;3(6):741.

Page 124: Preparation and characterization of biopolymer compounds

106

66. Berger E, Ramsay BA, Ramsay JA, Chavarie C, Braunegg G. PHB Recovery by hypochlorite

digestion of non-PHB biomass. Biotechnol Tech 1989;3(4):227-232.

67. Ramsay J, Berger E, Voyer R, Chavarie C, Ramsay B. Extraction of poly-3-hydroxybutyrate

using chlorinated solvents. Biotechnol Tech 1994;8(8):589-594.

68. Shozui F, Matsumoto K, Sasaki T, Taguchi S. Engineering of polyhydroxyalkanoate synthase

by Ser477X/Gln481X saturation mutagenesis for efficient production of 3-hydroxybutyrate-

based copolyesters. Appl Microbiol Biotechnol 2009;84(6):1117-1124.

69. Daly P, Bruce D, Melik D, Harrison G. Thermal degradation kinetics of poly(3-

hydroxybutyrate-co-3-hydroxyhexanoate). J Appl Polym Sci 2005;98(1):66-74.

70. Jiang X, Sun Z, Marchessault R, Ramsay J, Ramsay B. Biosynthesis and properties of medium-

chain-length polyhydroxyalkanoate with enriched content of the dominant monomer.

Biomacromolecules 2012;13(9):2926-2932

71. Sun Z, Ramsay J, Guay M, Ramsay B. Enhanced yield of medium-chain-length

polyhydroxyalkanoates from nonanoic acid by co-feeding glucose in carbon-limited, fed-batch

culture. J Biotechnol 2009;143(4):262-267.

72. Jiang X, Ramsay JA, Ramsay BA. Acetone extraction of mcl-PHA from Pseudomonas putida

KT2440. J Microbiol Methods 2006;67(2):212-219.

73. Dupont A, Harrison G. Conformation and dn/dc determination of cellulose in N,N-

dimethylacetamide containing lithium chloride. Carbohydr Polym 2004;58(3):233-243.

74. Barham P, Keller A, Otun E, Holmes P. Crystallization and morphology of a bacterial

thermoplastic - poly-3-hydroxybutyrate. J Mater Sci 1984;19(9):2781-2794.

Page 125: Preparation and characterization of biopolymer compounds

107

75. Beuermann S, Buback M, Davis T, Gilbert R, Hutchinson R, Kajiwara A, Klumperman B,

Russell G. Critically evaluated rate coefficients for free-radical polymerization, 3 - Propagation

rate coefficients for alkyl methacrylates. Macromol Chem Phys 2000;201(12):1355-1364.

76. Marchessault R, Monasterios C, Morin F, Sundararajan P. Chiral poly(beta-

hydroxyalkanoates) - an adaptable helix influenced by the alkane side-chain. Int J Biol

Macromol 1990;12(2):158-165.

77. Kim DY, Kim HW, Chung MG, Rhee YH. Biosynthesis, modification, and biodegradation of

bacterial medium-chain-length polyhydroxyalkanoates. J Microbiol 2007;45(2):87-97.

78. Ma P, Hristova-Bogaerds DG, Lemstra PJ, Zhang Y, Wang S. Toughening of PHBV/PBS and

PHB/PBS blends via in situ compatibilization using dicumyl peroxide as a free-radical grafting

initiator. Macromol Mater Eng 2012;297(5):402-410.

79. Ha CS, Cho WJ. Miscibility, properties, and biodegradability of microbial polyester

containing blends. Prog Polym Sci 2002;27(4):759-809.

80. Yu L, Dean K, Li L. Polymer blends and composites from renewable resources. Prog Polym

Sci 2006;31(6):576-602.

81. Zhang M, Thomas NL. Preparation and properties of polyhydroxybutyrate blended with

different types of starch. J Appl Polym Sci 2010;116(2):688-694.

82. Zhang M, Thomas NL. Blending polylactic acid with polyhydroxybutyrate: The effect on

thermal, mechanical, and biodegradation properties. Adv Polym Technol 2011;30(2):67-79.

83. Dufresne A, Vincendon M. Poly(3-hydroxybutyrate) and poly(3-hydroxyoctanoate) blends:

Morphology and mechanical behavior. Macromolecules 2000;33(8):2998-3008.

Page 126: Preparation and characterization of biopolymer compounds

108

84. Martelli SM, Sabirova J, Fakhoury FM, Dyzma A, de Meyer B, Soetaert W. Obtention and

characterization of poly(3-hydroxybutyricacid-co-hydroxyvaleric acid)/mcl-PHA based blends.

Fod Sci Technol-Leb 2012;47(2):386-392.

85. Nerkar M, Ramsay J, Ramsay B, Kontopoulou M, Hutchinson R. Determination of Mark-

Houwink Parameters and absolute molecular weight of medium-chain-length poly(3-

hydroxyalkanoates). J Polym Environ 2013;21(1):24-29.

86. Sin MC, Gan SN, Annuar MSM, Tan IKP. Thermodegradation of medium-chain-length poly(3-

hydroxyalkanoates) produced by Pseudomonas putida from oleic acid. Polym Degrad Stab

2010;95(12):2334-2342.

87. Pearce R, Marchessault R. Melting and crystallization in bacterial poly(beta-

hydroxyvalerate), PHV, and Blends with Poly(beta-hydroxybutyrate-co-hydroxyvalerate).

Macromolecules 1994;27(14):3869-3874.

88. Owen A, Heinzel J, Skrbic Z, Divjakovic V. Crystallization and melting behavior of PHB and

PHB/HV copolymer. Polymer 1992;33(7):1563-1567.

89. Organ S, Barham P. On the equilibrium melting temperature of Polyhydroxybutyrate.

Polymer 1993;34(10):2169-2174.

90. Gagnon K, Lenz R, Farris R, Fuller R. Chemical modification of bacterial elastomers .1.

peroxide cross-linking. Polymer 1994;35(20):4358-4367.

91. Dufresne A, Reche L, Marchessault R, Lacroix M. Gamma-ray crosslinking of poly (3-

hydroxyoctanoate-co-undecenoate). Int J Biol Macromol 2001;29(2):73-82.

Page 127: Preparation and characterization of biopolymer compounds

109

92. Park SJ, Kim TW, Kim MK, Lee SY, Lim S. Advanced bacterial polyhydroxyalkanoates:

Towards a versatile and sustainable platform for unnatural tailor-made polyesters. Biotechnol

Adv 2012;30(6):1196-1206.

93. Platel RH, Hodgson LM, Williams CK. Biocompatible initiators for lactide polymerization.

Polym Rev 2008;48(1):11-63.

94. Palade L, Lehermeier H, Dorgan J. Melt rheology of high L-content poly(lactic acid).

Macromolecules 2001;34(5):1384-1390.

95. Dorgan J, Williams J, Lewis D. Melt rheology of poly(lactic acid): entanglement and chain

architecture effects. J Rheol 1999;43(5):1141-1155.

96. Othman N, Acosta-Ramirez A, Mehrkhodavandi P, Dorgan JR, Hatzikiriakos SG. Solution and

melt viscoelastic properties of controlled microstructure poly(lactide). J Rheol 2011;55(5):987-

1005.

97. Othman N, Xu C, Mehrkhodavandi P, Hatzikiriakos SG. Thermorheological and mechanical

behavior of polylactide and its enantiomeric diblock copolymers and blends. Polymer

2012;53(12):2443-2452.

98. Othman N, Jazrawi B, Mehrkhodavandi P, Hatzikiriakos SG. Wall slip and melt fracture of

poly(lactides). Rheol Acta 2012;51(4):357-369.

99. Liu J, Lou L, Yu W, Liao R, Li R, Zhou C. Long chain branching polylactide: structures and

properties. Polymer 2010;51(22):5186-5197.

100. Pilla S, Kim SG, Auer GK, Gong S, Park CB. Microcellular extrusion-foaming of polylactide

with chain extender. Polym Eng Sci 2009;49(8):1653-1660.

Page 128: Preparation and characterization of biopolymer compounds

110

101. Yu L, Toikka G, Dean K, Bateman S, Yuan Q, Filippou C, Tri Nguyen. Foaming behaviour and

cell structure of poly(lactic acid) after various modifications. Polym Int 2013;62(5):759-765.

102. Meng Q, Heuzey M, Carreau PJ. Effects of a multifunctional polymeric chain extender on

the properties of polylactide and polylactide/clay nanocomposites. Int Polym Proc

2012;27(5):505-516.

103. Najafi N, Heuzey MC, Carreau PJ. Crystallization behavior and morphology of polylactide

and PLA/clay nanocomposites in the presence of chain extenders. Polym Eng Sci

2013;53(5):1053-1064.

104. Maharana T, Mohanty B, Negi YS. Melt-solid polycondensation of lactic acid and its

biodegradability. Prog Polym Sci 2009;34(1):99-124.

105. Graebling D. Synthesis of branched polypropylene by a reactive extrusion process.

Macromolecules 2002;35(12):4602-4610.

106. Passaglia E, Coiai S, Augier S. Control of macromolecular architecture during the reactive

functionalization in the melt of olefin polymers. Prog Polym Sci 2009;34(9):911-947.

107. Yang S, Wu Z, Yang W, Yang M. Thermal and mechanical properties of chemical crosslinked

polylactide (PLA). Polym Test 2008;27(8):957-963.

108. You J, Lou L, Yu W, Zhou C. The preparation and crystallization of long chain branching

polylactide made by melt radicals reaction. J Appl Polym Sci 2013;129(4):1959-1970.

109. Fukada E. Poiseuille Medal Award Lecture: Piezoelectricity of biopolymers. Biorheology

1995;32(6):593-609.

110. Long Y, Shanks R, Stachurski Z. Kinetics of polymer crystallization. Prog Polym Sci

1995;20(4):651-701.

Page 129: Preparation and characterization of biopolymer compounds

111

111. El Mabrouk K, Parent JS, Chaudhary BI, Cong R. Chemical modification of PP architecture:

strategies for introducing long-chain branching. Polymer 2009;50(23):5390-5397.

112. Wu W, Parent JS, Sengupta SS, Chaudhary BI. Preparation of crosslinked microspheres and

porous solids from hydrocarbon solutions: a new variation of precipitation polymerization

chemistry. J Polym Sci Pol Chem 2009;47(23):6561-6570.

113. Tian J, Yu W, Zhou C. Crystallization kinetics of linear and long-chain branched

polypropylene. J Macromol Sci Phys 2006;45(5):969-985.

114. Anderson KS, Schreck KM, Hillmyer MA. Toughening polylactide. Polym Rev 2008;48(1):85-

108.

115. Verhoogt H, Ramsay BA, Favis BD. Polymer blends containing poly(3-hydroxyalkanoate)s.

Polymer 1994;35(24):5155-5169.

116. Imre B, Pukanszky B. Compatibilization in bio-based and biodegradable polymer blends.

Eur Polym J 2013;49(6):1215-1233.

117. Schreck KM, Hillmyer MA. Block copolymers and melt blends of polylactide with Nodax

(TM) microbial polyesters: Preparation and mechanical properties. J Biotechnol

2007;132(3):287-295.

118. Noda I, Satkowski M, Dowrey A, Marcott C. Polymer alloys of Nodax copolymers and

poly(lactic acid). Macromol Biosci 2004;4(3):269-275.

119. Ma P, Spoelstra AB, Schmit P, Lemstra PJ. Toughening of poly (lactic acid) by poly (beta-

hydroxybutyrate-co-beta-hydroxyvalerate) with high beta-hydroxyvalerate content. Eur Polym J

2013;49(6):1523-1531.

Page 130: Preparation and characterization of biopolymer compounds

112

120. Bhardwaj R, Mohanty AK. Modification of brittle polylactide by novel hyperbranched

polymer-based nanostructures. Biomacromolecules 2007;8(8):2476-2484.

121. Nyambo C, Misra M, Mohanty AK. Toughening of brittle poly(lactide) with hyperbranched

poly(ester-amide) and isocyanate-terminated prepolymer of polybutadiene. J Mater Sci

2012;47(13):5158-5168.

122. Semba T, Kitagawa K, Ishiaku U, Hamada H. The effect of crosslinking on the mechanical

properties of polylactic acid/polycaprolactone blends. J Appl Polym Sci 2006;101(3):1816-1825.

123. Wang R, Wang S, Zhang Y, Wan C, Ma P. Toughening modification of PLLA/PBS blends via

in-situ compatibilization. Polym Eng Sci 2009;49(1):26-33.

124. Ma P, Cai X, Zhang Y, Wang S, Dong W, Chen M, Lemstra PJ. In-situ compatibilization of

poly(lactic acid) and poly(butylene adipate-co-terephthalate) blends by using dicumyl peroxide

as a free-radical initiator. Polym Degrad Stab 2014;102:145-151.

125. Dong W, Ma P, Wang S, Chen M, Cai X, Zhang Y. Effect of partial crosslinking on

morphology and properties of the poly(beta-hydroxybutyrate)/poly(D,L-lactic acid) blends.

Polym Degrad Stab 2013;98(9):1549-1555.

126. Nerkar M, Ramsay JA, Ramsay BA, Kontopoulou M, Hutchinson RA. Determination of

Mark-Houwink Parameters and Absolute Molecular Weight of Medium-Chain-Length Poly(3-

Hydroxyalkanoates). J Polym Environ 2013;21(1):24-29.

127. Ma J, La LTB, Zaman I, Meng Q, Luong L, Ogilvie D, Kuan H. Fabrication, structure and

properties of epoxy/metal nanocomposites. Macromol Mater Eng 2011;296(5):465-474.

128. Abdousabets S, Patel R. Morphology of elastomeric alloys. Rubber Chem Technol

1991;64(5):769-779.

Page 131: Preparation and characterization of biopolymer compounds

113

129. Palierne J. Linear Rheology of viscoelastic emulsions with Interfacial tension. Rheologica

Acta 1990;29(3):204-214.

130. Wu S. Formation of dispersed phase in incompatible polymer blends - interfacial and

rheological effects. Polym Eng Sci 1987;27(5):335-343.

Page 132: Preparation and characterization of biopolymer compounds

114

Appendix A - Improved viscosity ratio and compatibility of poly (lactic

acid) and polyhydroxyoctanoate blends

Introduction

In Chapter 6 it was reported that the reason for the very coarse morphology of poly (lactic acid)

(PLA)/polyhydroxyoctanoate (PHO) blend is the significant viscosity mismatch between the

blend components. This is evident from Figure A.1, which summarizes the complex viscosities of

the materials under consideration. Both polymers had Newtonian behavior, with viscosities of

661 Pa.s and 13 Pa.s for PLA and PHO respectively, at 180°C, resulting in a viscosity ratio

(defined as the ratio of the viscosity of the dispersed phase over the viscosity of the matrix) of

0.02.

The shortcoming of viscosity mismatch between PHO and poly-3-hydroxybutyrate (PHB) was

addressed in chapter 4 by free radical mediated cross-linking using peroxide. In this appendix,

the same approach is followed to increase the viscosity of PHO and reduce the viscosity

mismatch with PLA. Additionally epoxidation of PHO was explored to improve its compatibility

with PLA by reaction of epoxy groups with hydroxyl and carboxyl groups of PLA. Furthermore it

was anticipated that the epoxy groups would act as a chain extender to improve the viscosity of

PHO [1].

Page 133: Preparation and characterization of biopolymer compounds

115

Experimental

Cross-linking of PHO

PHO was crosslinked as described in Chapter 4 (Section 4.2.3) and dry mixed with PLA before

feeding in the DSM micro compounder. The gel content of the peroxide cross-linked MCL PHA

was measured as described in chapter 4.

Epoxidation of PHO

10-epoxyundecanoic acid was prepared as described below and was used to prepare epoxidized

PHO (ePHO). 50 g of 10-undecenoic acid were dissolved in 25 ml of anhydrous dichloromethane

and the solution was placed in an ice bath (0 oC). m-Chloroperbenzoic acid (mCPBA) was

purified by washing with a phosphate buffer solution of pH 7.5 and dried under reduced

pressure at room temperature. 65 g of purified mCPBA dissolved in 450 ml of anhydrous

dichloromethane were added to the 10-undecenoic solution drop wise, under continuous

stirring with a magnetic stirrer in an ice bath. The solution was then stirred for 24h until a

white precipitate formed. After filtration to eliminate the m-chlorobenzoic acid the filtrate was

washed with a 0.80 M solution of sodium sulphite until peroxide was no longer detectable with

peroxide test paper. The organic layer was finally washed with distilled water until pH test

paper indicated that the washings were neutral. The solution was then dried under reduced

pressure to yield a dry white powder.

Nuclear magnetic resonance characterization

Proton (1H) and carbon (13C) NMR spectra were recorded on a Bruker DPX300 NMR

spectrometer at 300 MHz in deuterated chloroform (CDCl3) as solvent. The chemical shifts

(ppm) for 1H and 13C NMR were referenced relative to tetramethylsilane (TMS, 0.00 ppm) as the

Page 134: Preparation and characterization of biopolymer compounds

116

internal reference. The solution concentration was 0.5 w/v% (20 mg of sample in the NMR

tube) and the solution was filtered prior to placing it in the NMR tube.

Yield = 99%; mp = 39 oC. 1H-NMR (300 MHz, CDCl3, δ ppm); 1.9-1.3 (m, 12H, (CH2)6; 2.3 (t, 2H,

CH2-CO2H); 2.5 (dd, 1H, CH epoxy); 2.8 (t, 1H, Ha of CH2 epoxy); 2.9 (d, 1H, Hb of CH2 epoxy).

Results and Discussion

Cross-linked PHO (xPHO)

Complex viscosities of the xPHO, are summarized in Figure A.1. Cross-linking led to significant

increase in shear thinning behavior and loss of the Newtonian plateau as expected for cross-

linked polymer with high cross-linked densities.

Given the change in the viscosity-shear rate dependence, xPHO can only match the viscosity of

PLA in a very narrow frequency/shear rate range. As shown in Figure A.1, lower peroxide

loading (0.06 wt.%) matched the viscosity of PLA at lower frequency region. At higher peroxide

level, 0.2 wt.%, the viscosity of PHO was equivalent to PLA viscosity in the mid frequency

region. In order to match the viscosity of PHO at high shear rates of about 100 s-1 (equivalent to

shear rate in processing), 0.5 wt.% of peroxide was needed, hence PHO was cross-linked using

0.5 wt.% peroxide and blended with PLA for further evaluations.

Page 135: Preparation and characterization of biopolymer compounds

117

Figure A.1 Complex viscosity of PLA, PHO, xPHO and ePHO at 180°C

The mechanical properties of PLA blends containing xPHO and ePHO are summarized in Table

A.1. xPHO matched the viscosity of PLA in the high frequency region and had pronounced shear

thinning characteristics, however, it did not show any improvement in mechanical properties as

the elastic nature of PHO was altered due to the high gel content (97%) of xPHO.

1 10 100

Co

mp

lex V

isco

sit

y (

Pa s

)

Frequency (rad/s)

PLA PHO 0.2 wt % Peroxide

0.06 wt % Peroxide 0.5 wt % Peroxide ePHO

100

101

102

103

104

105

Page 136: Preparation and characterization of biopolymer compounds

118

Table A.1 Mechanical properties of PLA and blends containing PHO, ePHO and xPHO

Epoxidized PHO

The compositions of the prepared 10-epoxyundecanoic acid and the ePHO were determined by

1H NMR spectroscopy. The spectra were consistent with the expected structures. The

epoxidation reaction of 10-undecanoic acid was followed by 1H-NMR. During the epoxidation

the characteristic signals corresponding to the unsaturated side group (2.0 ppm –CH2-, 4.9 ppm

=CH2, 5.8 ppm –CH=) disappear and are replaced by peaks associated with the oxirane group at

2.5 ppm (c, 0-CH-, multiplet) and 2.75-2.9 ppm (a and b, -O-CH2, triplet and multiplet,

respectively), while the –CH2- group is now found at 1.3 ppm (e). The 1H NMR spectrum of

ePHA, together with the corresponding chemical shift assignments, is presented in Figure A.2.

The percentage of epoxy groups in the prepared ePHA was calculated by comparing the oxirane

signals (i and j in particular, due to the fact the h is overlapped by the CH2 side chain groups)

with the -CH3 signal of the unmodified side chain (f). The content of epoxy modified monomer

in the resulting ePHO was 12 mole %, as determined by the 1H NMR.

Tensile

stress (MPa)

Tensile

Strain (%)

Unnotched

Impact (KJ/m2)

Young's Modulus

(MPa)

PLA 74 (±3) 14 (±1) 32 (±5) 670 (±74)

10 % PHO 50 (±4) 35 (±15) 65 (±5) 696 (±29)

10 % xPHO 40 (±2) 11 (±1) 23 (±1) 567 (±29)

10 % ePHO 47 (±4) 32 (±8) 57 (±11) 551 (±36)

Page 137: Preparation and characterization of biopolymer compounds

119

Figure A.2. NMR spectra of ePHA (12% mol/mol 10-epoxyundecanoic acid/unmodified

monomer)

Epoxidation of PHO did not show any increase in viscosity of PHO rather it was diminished

(Figure A.1).

Blends of PLA with ePHO

The compatibility of the blend components in immiscible blends is associated with their

interfacial tension. The Palierne model [2] was used to obtain an estimate of the interfacial

tension between PLA and PHO by fitting the complex modulus of the blend as a function of

frequency (A.3).

a

c

d

g

e

f

jh

b

c

d

b

i

a

b, h

f

d, e

c, g

i j

ji

d, h

a

Meth

an

ol

Ac

eto

ne

Chemical Shift ( )

Page 138: Preparation and characterization of biopolymer compounds

120

The Palierne model relates viscoelasticity of emulsion with the viscoelasticity of matrix and

dispersed phase, droplet size and droplet size distribution of dispersed phase and the

interphase surface tension of the blend components. It can be expressed as

( ) ( ) ( ( )

( )) (A.1)

H(ω) =

(

)

(

)

Gd* and Gm* are complex moduli of the dispersed phase and matrix respectively, α, the

interfacial tension, φ, the volume fraction of disperse phase, ω is the studied frequency and R is

the particle radius.

Immiscible blends of PLA and PHO formed droplet-matrix morphology with PLA as matrix phase

and PHO forming spherical dispersed phase. Scanning electron microscopy (SEM) with details

mentioned in chapter 4 was used to characterize blend morphology. Volume average diameter

of PHO domain determined using image analysis software SigmaScan Pro was 1.73 mm.

Interfacial tension was calculated by substituting volume average diameter and volume fraction

of PHO in equation A.1. The estimated interfacial tensions were 2.1 and 0.6 N/mm for PLA-PHO

and PLA-ePHO respectively. The epoxy groups of ePHO plausibly reacted with carboxylic acid

end group of PLA improving their compatibility and hence reducing interfacial tension.

Page 139: Preparation and characterization of biopolymer compounds

121

Figure A.3 Elastic modulus, G', of the PLA (matrix), PHO (droplet) and the 95/5 PLA/PHO blend

as a function of frequency at 180°C, and fit of the data using the Palierne model.

The mechanical properties of PLA/ePHO blend are depicted in Table A.1. Irrespective of

improved compatibility as a result of epoxidation, enhancement in properties was not

equivalent to that of neat PHO. ePHO that had viscosity lower than neat PHO, yielded a higher

viscosity mismatch and a coarser blend morphology (Figure A.4).

Figure A.4 Scanning electron microscopy of PLA blends containing 10 % of a) PHO, b) ePHO and

c) xPHO (0.5 wt.% of peroxide)

1.E-03

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

0.01 0.1 1 10 100

G' (P

a)

Frequency (rad/s)

Experimental G'

Palierne

Matrix G'

Droplet G'

20mm 20mm 20mm

a) b) c)

Page 140: Preparation and characterization of biopolymer compounds

122

To further identify the mechanism for the morphology development the droplet size of PHO

and ePHO in PLA matrix was predicted using Wu model [3] for viscoelastic liquids as shown in

equation A.2 and further it was compared with particle size obtained from SEM images.

The model correlates viscosity of blend components, interfacial tension, droplet size and shear

rate according to

(

)

(A.2)

where, – shear rate, Ƞm- viscosity of matrix, Ƞd- viscosity of dispersed, σ- surface tension, D-

diameter of particle (+ve sign for λ>1, -ve for λ<1)

The predicted particle size at a shear rate of 100 S-1 was 4mm which is fairly close to the particle

size obtained from SEM images. The particle size of ePHO was higher at 8mm

In spite of the lower interfacial tension in ePHO/PLA, blend morphology was coarser, indicating

that effect of viscosity mismatch was more prominent than the interfacial tension reduction.

This explains the lack of improvements in morphology and hence mechanical properties.

Conclusion

Epoxidation of PHO improved its compatibility with PLA, however did not show any

improvement in elongation and impact properties due to increased viscosity mismatch with

PLA. Cross-linking of PHO improved its viscosity substantially. Irrespective of lowering viscosity

mismatch, xPHO/PLA blend did not enhance ductility, mainly due to high gel content altering its

elastic nature.

Page 141: Preparation and characterization of biopolymer compounds

123

References

1. Al-Itry R, Lamnawar K, Maazouz A. Improvement of thermal stability, rheological and

mechanical properties of PLA, PBAT and their blends by reactive extrusion with functionalized

epoxy. Polym Degrad Stab 2012;97(10):1898-1914.

2. Palierne J. Linear rheology of viscoelastic emulsions with interfacial tension. Rheologica Acta

1990;29(3):204-214.

3. Wu S. Formation of dispersed phase in incompatible polymer blends - interfacial and

rheological effects. Polym Eng Sci 1987;27(5):335-343